This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Martingale measure associated with the critical 2d2d stochastic heat flow

Makoto Nakashima
Abstract

In [14], they proved the convergence of the finite dimensional time distribution of the rescaled random fields derived from the discrete stochastic heat equation of 2d2d-directed polymers in random environment in the critical window. The scaling limit is called critical 2d2d stochastic heat flow (SHF).

In this paper, we will show that the critical 2d2d SHF is a continuous semimartingale. Moreover, we will consider the martingale problem associated with the critical 2d2d SHF in a similar fashion to the super Brownian motion which is one of the well-known measure valued process. Also, we define the martingale measure associated with the critical 2d2d SHF in the sense of [45, Chapter 2].

The quadratic variation of the martingale measure gives information of the regularity of the critical 2d2d SHF.

MSC 2020 Subject Classification: Primary 60H17. Secondary 65C35, 60G44.

Key words: Critical 2d2d stochastic heat flow, Directed polymers in random environment, Martingale problem, Martingale measure.

1 Introduction and main results

Kardar-Parisi-Zhang considered the stochastic partial differential equations (KPZ equation) which describes the evolution of the random interface:

th=νΔh+λ2|h|2+𝒲˙,\displaystyle\partial_{t}h=\nu\Delta h+\frac{\lambda}{2}\left|\nabla h\right|^{2}+\dot{\mathcal{W}}, (KPZ)

where 𝒲˙\dot{\mathcal{W}} is space-time white noise on [0,)×d[0,\infty)\times\mathbb{R}^{d} [33]. It is ill-posed due to the term |h|2|\nabla h|^{2} which should be the square of distribution. In [33], they also considered the formal transformation u(t,x)=exp(λ2νh(t,x))u(t,x)=\exp\left(\frac{\lambda}{2\nu}h(t,x)\right) and obtained the multiplicative stochastic heat equation

tu=νΔu+λ2νu𝒲˙.\displaystyle\partial_{t}u=\nu\Delta u+\frac{\lambda}{2\nu}u\dot{\mathcal{W}}. (SHE)

In [3], Bertini and Giacomin studied this transformation mathematically in dimension 11 in the analysis of weakly asymmetric simple exclusion process and SOS process. They considered the mollified stochastic heat equation

tuε=12Δuελuε𝒲˙ε,\displaystyle\partial_{t}u^{\varepsilon}=\frac{1}{2}\Delta u^{\varepsilon}-\lambda u^{\varepsilon}\dot{\mathcal{W}}^{\varepsilon}, (SHEε\text{SHE}_{\varepsilon})

under suitable initial condition u0(x)>0u_{0}(x)>0, where 𝒲˙ε(t,x):=jε(xy)𝒲(t,y)˙dy\dot{\mathcal{W}}^{\varepsilon}(t,x):=\int_{{\mathbb{R}}}j_{\varepsilon}(x-y)\dot{\mathcal{W}(t,y)}\text{\rm d}y is a space-mollified noise for the probability density jCc()j\in C_{c}^{\infty}({\mathbb{R}}) with j(x)=j(x)j(x)=j(-x) and jε(x)=ε1j(xε)j^{\varepsilon}(x)=\varepsilon^{-1}j\left(\frac{x}{\varepsilon}\right). Then, hε(t,x):=loguε(t,x)h^{\varepsilon}(t,x):=\log u^{\varepsilon}(t,x) satisfies a mollified KPZ equation

thε=12Δhελ2(|hε|2Jε(0))+𝒲˙,\displaystyle\partial_{t}h^{\varepsilon}=\frac{1}{2}\Delta h^{\varepsilon}-\frac{\lambda}{2}\left(|\nabla h^{\varepsilon}|^{2}-J_{\varepsilon}(0)\right)+\dot{\mathcal{W}}, (KPZε)

where Cε(y)=1εj(x)j(yx)dyC_{\varepsilon}(y)=\frac{1}{\varepsilon}\int j(x)j(y-x)\text{\rm d}y.

They proved that uεu^{\varepsilon} converges a.s. and in LpL^{p} to the solution of (SHE) uniformly on the compact set in [0,)×[0,\infty)\times{\mathbb{R}}. Also, Mueller proved that if u0u_{0} is nonnegative, not identically zero, and continuous, then u(t,x)u(t,x) is strictly positive on {\mathbb{R}} for all t>0t>0 [40]. Thus, we find that hεh^{\varepsilon} also converges to some process 𝔥(t,x):=logu(t,x)\mathfrak{h}(t,x):=\log u(t,x) a.s., which is so called Cole-Hopf solution of (KPZ).

For KPZ equation, Hairer developed the regularity structures and showed the existence of the distributional solution to (KPZ)[27, 28, 19]. In the singular SPDE literature, the existence of the solution of (KPZ) has been proved by Gubinelli-Imkeller-Perkowski via the Paracontrolled calculus[23], by Gonçalves-Jara via the energy solutions[24], and by Kupiainen via the renormalization group approach[34].

Many attempts have been made to construct solutions of (KPZ) for d2d\geq 2, where the dimension d=1d=1 (d=2d=2, d3d\geq 3) is called sub-critical(critical, super-critical) in the literature of singular SPDE[28]. Also, d=1d=1 and d3d\geq 3 are called ultraviolet superrenormalizable and infrared renormalizable in the physicists’ language [39], respectively.

One approach is to modify the Bertini-Giacomin’s idea: We consider the stochastic heat equation

tuβ0,ε=12Δuβ0,εβεuβ0,ε𝒲˙,\displaystyle\partial_{t}u^{\beta_{0},\varepsilon}=\frac{1}{2}\Delta u^{\beta_{0},\varepsilon}-\beta_{\varepsilon}u^{\beta_{0},\varepsilon}\dot{\mathcal{W}}, (SHEβ0,ε{}_{\beta_{0},\varepsilon})

where β00\beta_{0}\geq 0 and βε\beta_{\varepsilon} is defined by

βε={β02πlogεd=2β0εd21d3.\displaystyle\beta_{\varepsilon}=\begin{cases}\beta_{0}\sqrt{\frac{2\pi}{-\log\varepsilon}}\quad&d=2\\ \beta_{0}\varepsilon^{\frac{d}{2}-1}\quad&d\geq 3.\end{cases}

In (SHEβ0,ε{}_{\beta_{0},\varepsilon}), the strength of the noise term of (SHEε\text{SHE}_{\varepsilon}) is tuned. Then, it has been proved that there exists a phase transition of the one-point distribution in the following sense. There exists βc>0\beta_{c}>0 such that if 0β0<βc0\leq\beta_{0}<\beta_{c} (subcriticale regime), then uβ0,ε(t,x)u^{\beta_{0},\varepsilon}(t,x) converges to a non-trivial random variable 𝔲β0,(t,x)\mathfrak{u}^{\beta_{0},\star}(t,x) for each t>0t>0 and xdx\in{\mathbb{R}}^{d} as ε0\varepsilon\to 0, and if β0>βc\beta_{0}>\beta_{c} (supercritical regime), then uβ0,ε(t,x)u^{\beta_{0},\varepsilon}(t,x) vanishes[38, 10]. However, we know that if β0<βc\beta_{0}<\beta_{c}, then uβ0,εu^{\beta_{0},\varepsilon} converges to the solution of the non-random heat equation in the weak sense, i.e. for each test function ϕCc\phi\in C_{c}, duβ0,ε(t,x)ϕ(x)dxd×du0(y)pt(x,y)ϕ(y)dxdy\int_{{\mathbb{R}}^{d}}u^{\beta_{0},\varepsilon}(t,x)\phi(x)\text{\rm d}x\to\int_{{\mathbb{R}}^{d}\times{\mathbb{R}}^{d}}u_{0}(y)p_{t}(x,y)\phi(y)\text{\rm d}x\text{\rm d}y in probability, where pt(x,y)p_{t}(x,y) is the Gaussian density with mean 0 and variance tt[38, 9, 10], which can be regarded as the law of large numbers. Thus, we cannot give the new definition of (SHE) (and hence (KPZ)) for the subcritical regime.

Remark 1.1.

We know that in the subcritical regime, the fluctuations of the centered field uβ0,εu^{\beta_{0},\varepsilon} and hβ0,ε=loguβ0,εh^{\beta_{0},\varepsilon}=\log u^{\beta_{0},\varepsilon} converges to the solution of Edwards-Wilkinson type equation [26, 17, 26, 39, 13, 6, 7, 8, 9, 36, 5] in so-called L2L^{2}-regime, which can be regarded as the central limit theorem. For d3d\geq 3, Junk and Nakajima discussed the fluctuation of the centered field of directed polymers (the discrete counterpart of (SHEε\text{SHE}_{\varepsilon})) beyond the L2L^{2}-regime[31, 32, 29].

One is interested in the critical case β0=βc=1\beta_{0}=\beta_{c}=1 for d=2d=2 is the interesting phase. In [2], Bertini and Cancrini retake βe\beta_{e} and consider the critical window around the critical point βc=1\beta_{c}=1 given by

βε=2πlogε+ρ+o(1)(logε)2,ρ.\displaystyle\beta_{\varepsilon}=\sqrt{\frac{2\pi}{-\log\varepsilon}+\frac{\rho+o(1)}{\left(-\log\varepsilon\right)^{2}}},\quad\rho\in{\mathbb{R}}.

They proved that if u0,ψL2(2)u_{0},\psi\in L^{2}({\mathbb{R}}^{2}), then the variance of 2uβ0,ε(t,x)ψ(x)dx\int_{{\mathbb{R}}^{2}}u^{\beta_{0},\varepsilon}(t,x)\psi(x)\text{\rm d}x converges to the nontrivial quantity which has the same form as (2.13). Thus, the tightness of the random field uβ0,ε(t,x)u^{\beta_{0},\varepsilon}(t,x) follows. Moreover, the finiteness of the higher moments has been verified by Caravenna-Sun-Zygouras[12] and Gu-Quastel-Tsai[25] so that the limit point filed should be random field. Finally, the weak convergence of the random field was proved by Caravenna-Sun-Zygouras for directed polymers setting in finite dimensional time distribution sense in [14] and by Tsai for (SHEβ0,ε{}_{\beta_{0},\varepsilon}) in process level in [44]. Thus, the limit 𝒵\mathscr{Z} can be regarded as the solution to (SHE) for d=2d=2.

Our main results concern the martingale part of (SHE) for 𝒵\mathscr{Z}. If we rewrite (SHE) formally by

2𝒵(t,x)ψ(x)dx2𝒵(0,x)ϕ(x)ψ(x)dx\displaystyle\int_{{\mathbb{R}}^{2}}\mathscr{Z}(t,x)\psi(x)\text{\rm d}x-\int_{{\mathbb{R}}^{2}}\mathscr{Z}(0,x)\phi(x)\psi(x)\text{\rm d}x
=0t212Δ𝒵(s,x)ψ(x)dxds+0t2β𝒵(s,x)ψ(x)𝒲˙(ds,dx),\displaystyle=\int_{0}^{t}\int_{{\mathbb{R}}^{2}}\frac{1}{2}\Delta\mathscr{Z}(s,x)\psi(x)\text{\rm d}x\text{\rm d}s+\int_{0}^{t}\int_{{\mathbb{R}}^{2}}\beta\mathscr{Z}(s,x)\psi(x)\dot{\mathcal{W}}(\text{\rm d}s,\text{\rm d}x),

then we may believe the stochastic integral of the last term would be martingale. However, [15] shows that the random field 𝖹t(dx)\mathsf{Z}_{t}(\text{\rm d}x) is singular with respect to Lebesgue measure. Therefore, the stochastic integral is formal. In our main results, we will show that it is indeed a martingale and its quadratic variation can be described in terms of 𝖹\mathsf{Z}.

1.1 Setting and known results

In this paper, we consider the model in the discrete setting as in [14].

Let {ωn,x}n,x2\{\omega_{n,x}\}_{n\in{\mathbb{Z}},x\in{\mathbb{Z}}^{2}} be i.i.d. random variables with the law \mathbb{P} such that

𝔼[ωn,x]=0,𝔼[ωn,x2]=1,λ(β):=log𝔼[eβωn,x]<for small β>0.\displaystyle\mathbb{E}[\omega_{n,x}]=0,\quad\mathbb{E}[\omega_{n,x}^{2}]=1,\quad\lambda(\beta):=\log\mathbb{E}\left[e^{\beta\omega_{n,x}}\right]<\infty\quad\text{for small }\beta>0. (1.1)

Now, we introduce the random fields according to [14, 15].

Let {Sn}n\{S_{n}\}_{n\in{\mathbb{Z}}} be an irreducible, symmetric, and aperiodic random walk on 2{\mathbb{Z}}^{2} whose increment ξ:=S1S0\xi:=S_{1}-S_{0} has mean 0 and covariance matrix being the identity matrix II. Let PP and EE denote probability and expectation for SS. Also, we assume that ξ\xi has finite support, i.e. there exists a finite set Σ2\Sigma\subset{\mathbb{Z}}^{2} such that P(ξΣ)=1P(\xi\in\Sigma)=1.

We define the point-to-point partition function of 2d2d-directed polymers in random environment

ZM,Nβ(x,y):=E[exp(i=M+1N(βωi,Siλ(β)))𝟙SN=y|SM=x]\displaystyle Z_{M,N}^{\beta}(x,y):=E\left[\left.\exp\left(\sum_{i=M+1}^{N}\left(\beta\omega_{i,S_{i}}-\lambda(\beta)\right)\right)\mathbbm{1}_{S_{N}=y}\right|S_{M}=x\right] (1.2)
Z¯M,Nβ(x,y):=E[exp(i=M+1N1(βωi,Siλ(β)))𝟙SN=y|SM=x]\displaystyle\overline{Z}_{M,N}^{\beta}(x,y):=E\left[\left.\exp\left(\sum_{i=M+1}^{N-1}\left(\beta\omega_{i,S_{i}}-\lambda(\beta)\right)\right)\mathbbm{1}_{S_{N}=y}\right|S_{M}=x\right] (1.3)

for M,NM,N\in{\mathbb{Z}} (MNM\leq N) and x,y2x,y\in{\mathbb{Z}}^{2}, β0\beta\geq 0, where we use the convention n=M+1k{}=:0\sum_{n=M+1}^{k}\{\dots\}=:0 for k<M+1k<M+1.

We note that

𝔼[ZM,Nβ(x,y)]=𝔼[Z¯M,Nβ(x,y)]=P(SN=y|SM=x)=qNM(yx).\displaystyle\mathbb{E}\left[Z_{M,N}^{\beta}(x,y)\right]=\mathbb{E}\left[\overline{Z}_{M,N}^{\beta}(x,y)\right]=P(S_{N}=y|S_{M}=x)=q_{N-M}(y-x).

where we denote by qn(x)q_{n}(x) the transition probability kernel of the underlying random walk starting at 0 i.e.

qn(x):=P(Sn=x|S0=0)\displaystyle q_{n}(x):=P(S_{n}=x|S_{0}=0)

for x2x\in{\mathbb{Z}}^{2}, n0n\geq 0.

For ss\in{\mathbb{R}}, we denote by s\left\lfloor{s}\right\rfloor the greatest integer nsn\leq s. For x2x\in{\mathbb{R}}^{2}, we define x\lfloor x\rfloor by the closest point z2z\in{\mathbb{Z}}^{2} of x2x\in{\mathbb{R}}^{2} (if more than two points exist, we choose the smallest one in the lexicographic order). Also, we write s=s\left\lfloor{s}\right\rfloor=s and x=x\left\lfloor{x}\right\rfloor=x for ss\in{\mathbb{R}} and x2x\in{\mathbb{R}}^{2} if it is clear from the context that ss and xx should be an integer and a lattice point.

For fixed N1N\geq 1, we focused on rescaled random measure valued flows which are defined by

𝖹N;s,tβ:=1NZNs,Ntβ(x,y)δxNδyN,\displaystyle\mathsf{Z}^{\beta}_{N;s,t}:=\frac{1}{N}\sum Z^{\beta}_{\left\lfloor{Ns}\right\rfloor,\left\lfloor{Nt}\right\rfloor}\left(x,y\right)\delta_{\frac{x}{\sqrt{N}}}\delta_{\frac{y}{\sqrt{N}}},

for <st<-\infty<s\leq t<\infty, where we take the summation over all x,y2x,y\in{\mathbb{Z}}^{2}. Then, we have for ϕCc2(2)\phi\in C_{c}^{2}({\mathbb{R}}^{2}), ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}), and <st<-\infty<s\leq t<\infty

𝖹N;s,tβ(ϕ,ψ)\displaystyle\mathsf{Z}_{N;s,t}^{\beta}(\phi,\psi)
:=1Nx2y2ϕ(xN)ZNs,Ntβ(x,y)ψ(yN)\displaystyle:=\frac{1}{N}\sum_{x\in{\mathbb{Z}}^{2}}\sum_{y\in{\mathbb{Z}}^{2}}\phi\left(\frac{x}{\sqrt{N}}\right)Z_{\left\lfloor{Ns}\right\rfloor,\left\lfloor{Nt}\right\rfloor}^{\beta}(x,y)\psi\left(\frac{y}{\sqrt{N}}\right)
=1Ny2E[ϕ(SNtN)exp(i=Ns+1Nt(βωi,SNti+Nsλ(β)))|SNs=y]ψ(yN)\displaystyle=\frac{1}{N}\sum_{y\in{\mathbb{Z}}^{2}}E\left[\left.\phi\left(\frac{S_{\left\lfloor{Nt}\right\rfloor}}{\sqrt{N}}\right)\exp\left(\sum_{i=\left\lfloor{Ns}\right\rfloor+1}^{\left\lfloor{Nt}\right\rfloor}\left(\beta\omega_{i,S_{\left\lfloor{Nt}\right\rfloor-i+\left\lfloor{Ns}\right\rfloor}}-\lambda(\beta)\right)\right)\right|S_{\left\lfloor{Ns}\right\rfloor}=y\right]\psi\left(\frac{y}{\sqrt{N}}\right) (1.4)

We define 𝖹¯N;s,tβ\overline{\mathsf{Z}}_{N;s,t}^{\beta} and 𝖹¯N;s,tβ(ϕ,ψ)\overline{\mathsf{Z}}_{N;s,t}^{\beta}(\phi,\psi) by replacing ZβZ^{\beta} by Z¯β\overline{Z}^{\beta}.

We equip with the space of locally finite measures on 2×2{\mathbb{R}}^{2}\times{\mathbb{R}}^{2} and finite measures on 2{\mathbb{R}}^{2} with the topology of vague convergence and the topology of weak convergence, respectively:

μNμdeffor any ϕCc(2×2),ϕ(x,y)μN(dx,dy)ϕ(x,y)μ(dx,dy)\displaystyle\mu_{N}\to\mu\overset{\text{def}}{\Leftrightarrow}\text{for any $\phi\in C_{c}({\mathbb{R}}^{2}\times{\mathbb{R}}^{2})$},\int\phi(x,y)\mu_{N}(\text{\rm d}x,\text{\rm d}y)\to\int\phi(x,y)\mu(\text{\rm d}x,\text{\rm d}y)

for μN\mu_{N} and μ\mu the locally finite measures on 2×2{\mathbb{R}}^{2}\times{\mathbb{R}}^{2} and

νNνdeffor any ϕCb(2),ϕ(x)νN(dx)ϕ(x)ν(dx)\displaystyle\nu_{N}\to\nu\overset{\text{def}}{\Leftrightarrow}\text{for any $\phi\in C_{b}({\mathbb{R}}^{2})$},\int\phi(x)\nu_{N}(\text{\rm d}x)\to\int\phi(x)\nu(\text{\rm d}x)

for νN\nu_{N} and ν\nu the finite measures on 2{\mathbb{R}}^{2}.

Remark 1.2.

The sequence of random measure-valued flow 𝖹Nβ={𝖹N;s,tβ}<st<\mathsf{Z}_{N}^{\beta}=\left\{\mathsf{Z}_{N;s,t}^{\beta}\right\}_{-\infty<s\leq t<\infty} are almost the same one considered in [14, 15]. The differences are as follows:

  1. (1)

    In [14, 15], they used 𝖹¯Nβ\overline{\mathsf{Z}}_{N}^{\beta} instead of 𝖹Nβ\mathsf{Z}_{N}^{\beta}.

  2. (2)

    In [14, 15], random measures are absolutely continuous with respect to Lebesgue measures by replacing Dirac measure by uniform measure on square.

However, these differences are negligible for the convergence.

To see the nontrivial limit of random measures obtained in [14, 15], we rescale the strength of disorders β\beta as β=βN\beta=\beta_{N} properly.

Let SS^{\prime} be an independent copy of SS and RNR_{N} be the expectation of the number of collisions of SS and SS^{\prime} up to time NN:

RN:=n=1NP(Sn=Sn)=n=1Nx2qn(x)2=n=1Nq2n(0).\displaystyle R_{N}:=\sum_{n=1}^{N}P(S_{n}=S_{n}^{\prime})=\sum_{n=1}^{N}\sum_{x\in{\mathbb{Z}}^{2}}q_{n}(x)^{2}=\sum_{n=1}^{N}q_{2n}(0).

Then, the local limit theorem below gives that the asymptotic behavior

RNlogN4π(1+o(1))\displaystyle R_{N}\sim\frac{\log N}{4\pi}(1+o(1)) (1.5)
Theorem 1.3.

(The local limit theorem [42, P7.9], [35, Theorem 2.3.5,Theorem 2.3.11]) Let qnq_{n} be the transition probability kernel defined as above. Then, we have

qn(x)=pn(x)+O(1n2)=pn(x)eO(1n+O(|x|4n3))\displaystyle q_{n}(x)=p_{n}(x)+O\left(\frac{1}{n^{2}}\right)=p_{n}(x)e^{O\left(\frac{1}{n}+O\left(\frac{|x|^{4}}{n^{3}}\right)\right)} (1.6)

for nn\in\mathbb{N}, x2x\in{\mathbb{Z}}^{2}, where

pt(x)=12πtexp(|x|22t)t>0,x2\displaystyle p_{t}(x)=\frac{1}{2\pi t}\exp\left(-\frac{|x|^{2}}{2t}\right)\qquad t>0,x\in{\mathbb{R}}^{2} (1.7)

is the Gaussian density on 2{\mathbb{R}}^{2} with mean 0 and variance tItI.

We choose the disorder strength β=βN\beta=\beta_{N} such that

σN2:=eλ(2βN)2λ(βN)1=1RN(1+ϑlogN(1+o(1)))as N\displaystyle\sigma_{N}^{2}:=e^{\lambda(2\beta_{N})-2\lambda(\beta_{N})}-1=\frac{1}{R_{N}}\left(1+\frac{\vartheta}{\log N}(1+o(1))\right)\quad\text{as $N\to\infty$} (1.8)

for some ϑ\vartheta\in{\mathbb{R}}.

Theorem 1.4.

[14, Theorem 1.1] [15, Theorem 6.1] The family of random measures 𝖹NβN={𝖹N;s,tβN}<st<\mathsf{Z}^{\beta_{N}}_{N}=\left\{\mathsf{Z}^{\beta_{N}}_{N;s,t}\right\}_{-\infty<s\leq t<\infty} converges in finite dimensional distribution to a unique limit (called Critical 2d2d Stochastic Heat Flow)

𝒵ϑ={𝒵s,tϑ(dx,dy)}<st<.\displaystyle\mathscr{Z}^{\vartheta}=\{\mathscr{Z}^{\vartheta}_{s,t}(\text{\rm d}x,\text{\rm d}y)\}_{-\infty<s\leq t<\infty}.

Moreover, the distribution of 𝒵ϑ\mathscr{Z}^{\vartheta} is independent of the choice of {ωn,x}n,x\{\omega_{n,x}\}_{n,x}.

Remark 1.5.

We will see that 𝔼[(𝖹N;s,tβN(ϕ,ψ)𝖹¯N;s,tβN(ϕ,ψ))2]0\mathbb{E}\left[\left(\mathsf{Z}^{\beta_{N}}_{N;s,t}(\phi,\psi)-\overline{\mathsf{Z}}^{\beta_{N}}_{N;s,t}(\phi,\psi)\right)^{2}\right]\to 0 in Remark 3.1, so {𝖹¯N;s,tβN}\left\{\overline{\mathsf{Z}}^{\beta_{N}}_{N;s,t}\right\} also converges to 𝒵θ\mathscr{Z}^{\theta}.

Remark 1.6.

The statement in Theorem 1.4 is a bit different from the original one in [14] at the point of the range of sts\leq t but it does not affect the proof in [14]. Also, [14, Theorem 9.1] says that for 0siti<0\leq s_{i}\leq t_{i}<\infty, ϕiCc(2)\phi_{i}\in C_{c}({\mathbb{R}}^{2}), and ψiCb(2)\psi_{i}\in C_{b}({\mathbb{R}}^{2}) (i=1,,ki=1,\dots,k)

{𝖹N;si,tiβN(ϕi,ψi)}i=1,,k{𝒵si,tiϑ(ϕi,ψi)}i=1,,k,\displaystyle\left\{\mathsf{Z}^{\beta_{N}}_{N;s_{i},t_{i}}(\phi_{i},\psi_{i})\right\}_{i=1,\dots,k}\Rightarrow\left\{\mathscr{Z}^{\vartheta}_{s_{i},t_{i}}(\phi_{i},\psi_{i})\right\}_{i=1,\dots,k},

where

𝒵s,tϑ(ϕ,ψ)\displaystyle\mathscr{Z}^{\vartheta}_{s,t}(\phi,\psi) =ϕ(x)ψ(y)𝒵s,tϑ(dx,dy).\displaystyle=\int\int\phi(x)\psi(y)\mathscr{Z}^{\vartheta}_{s,t}(\text{\rm d}x,\text{\rm d}y).
Remark 1.7.

In this paper, we focus only on the case s=0s=0 (so Ns=0\lfloor Ns\rfloor=0 ). Hence, we often omit the first coordinate of pair of times for a flow {Xs,t}<st<\{X_{s,t}\}_{-\infty<s\leq t<\infty}.

Remark 1.8.

In [44], Tsai proves the convergence of flows derived from the mollified stochastic heat equation (SHEβ0,ε{}_{\beta_{0},\varepsilon}) with the critical windows.

1.2 Measure valued process

We will show that for fixed ϕCc(d)\phi\in C_{c}({\mathbb{R}}^{d}),

𝒵ϑ(ϕ,dy):={𝒵tϑ(ϕ,dy)}t0\displaystyle\mathscr{Z}^{\vartheta}(\phi,\text{\rm d}y):=\left\{\mathscr{Z}^{\vartheta}_{t}(\phi,\text{\rm d}y)\right\}_{t\geq 0}

has a version which has continuous sample paths almost surely, where we define

𝒵tϑ,ϕ(dy)=𝒵tϑ(ϕ,dy):=ϕ(x)𝒵tϑ(dx,dy),\displaystyle\mathscr{Z}^{\vartheta,\phi}_{t}(\text{\rm d}y)=\mathscr{Z}^{\vartheta}_{t}(\phi,\text{\rm d}y):=\int\phi(x)\mathscr{Z}^{\vartheta}_{t}(\text{\rm d}x,\text{\rm d}y),

and give a semimartingale representation.

First of all, we recall some facts about measure-valued process from [41] which is a textbook of Dawson-Watanabe superdiffusions (or super-Brownian motions).

Let EE be a Polish space. Then, we define

C(E):=C(+,E)\displaystyle C(E):=C({\mathbb{R}}_{+},E) the set of EE-valued paths with the topology of uniform convergence on compacts,
D(E):=D(+,E)\displaystyle D(E):=D({\mathbb{R}}_{+},E) the set of càdlàg EE-valued paths with the Skorokhod J1J_{1}-topology.

Let MF(2)M_{F}({\mathbb{R}}^{2}) be the set of finite measures on 2{\mathbb{R}}^{2} with the topology of weak convergence. Then, MF(2)M_{F}({\mathbb{R}}^{2}) is also a Polish space [18, Theorem 3.1.7] and hence D(MF(2))D(M_{F}({\mathbb{R}}^{2})) is a Polish space.

For a càdlàg MF(2)M_{F}({\mathbb{R}}^{2})-valued process X={Xt}t0X=\{X_{t}\}_{t\geq 0}, we denote by

tX=u>tσ[Xr:0ru]\displaystyle{\mathcal{F}}_{t}^{X}=\bigcap_{u>t}\sigma[X_{r}:0\leq r\leq u] (1.9)

the right-continuous filtration generated by a process XX.

Our first main result shows the existence of a continuous version of 𝒵tϑ,ϕ\mathscr{Z}^{\vartheta,\phi}_{t}.

Theorem 1.9.

For each ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}), there exists a continuous MF(2)M_{F}({\mathbb{R}}^{2})-valued process 𝒵ϑ,ϕ={𝒵tϑ,ϕ}t0\mathscr{Z}^{\vartheta,\phi}=\{\mathscr{Z}_{t}^{\vartheta,\phi}\}_{t\geq 0} such that its finite dimensional distributions are identical to those of the Critical 2d2d SHF.

Remark 1.10.

In [44], Tsai obtains “another” critical 2d2d stochastic heat flows 𝒵~ϑ\widetilde{\mathscr{Z}}^{\vartheta^{\prime}} from the mollified stochastic heat equation (SHEβ0,ε{}_{\beta_{0},\varepsilon}) with the critical windows. We may expect that 𝒵ϑ\mathscr{Z}^{\vartheta} by [14] and 𝒵ϑ\mathscr{Z}^{\vartheta^{\prime}} by [44] are identical for some suitable pairs (ϑ,ϑ)(\vartheta,\vartheta^{\prime}), but it has not yet been verified.

In [44], he gives a characterization of the critical 2d2d SHF by four conditions, one of which is the conditions of continuity of the flows, i.e. the continuity in two parameters (s,t)(s,t). In Theorem 1.9, we have proved the continuity of the process, i.e. the continuity in one parameter tt.

Our second theorem gives a martingale problem of the measure-valued process 𝒵ϑ,ϕ\mathscr{Z}^{\vartheta,\phi}, which is similar to the form discussed in super-Brownian motion.

Theorem 1.11.

Let ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}), ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}), and ϑ\vartheta\in{\mathbb{R}}. Let 𝒵ϑ,ϕ(ψ):=ψ(x)𝒵ϑ,ϕ(dx)\mathscr{Z}^{\vartheta,\phi}(\psi):=\int\psi(x)\mathscr{Z}^{\vartheta,\phi}(\text{\rm d}x) be the continuous process. We define

tϑ,ϕ(ψ):=𝒵tϑ,ϕ(ψ)ϕ(x)ψ(y)dxdy0t𝒵sϑ,ϕ(12Δϕ)ds\displaystyle\mathscr{M}_{t}^{\vartheta,\phi}(\psi):=\mathscr{Z}^{\vartheta,\phi}_{t}(\psi)-\int\phi(x)\psi(y)\text{\rm d}x\text{\rm d}y-\int_{0}^{t}\mathscr{Z}^{\vartheta,\phi}_{s}\left(\frac{1}{2}\Delta\phi\right)\text{\rm d}s (1.10)

for t0t\geq 0. Then, tϑ,ϕ(ψ)\mathscr{M}_{t}^{\vartheta,\phi}(\psi) is a continuous {t𝒵ϑ}t0\left\{{\mathcal{F}}_{t}^{\mathscr{Z}^{\vartheta}}\right\}_{t\geq 0}-martingale such that

0ϑ,ϕ(ψ)=0\displaystyle\mathscr{M}_{0}^{\vartheta,\phi}(\psi)=0
ϑ,ϕ(ψ)t=limε04πlogε0t2(𝒵uϑ,ϕ(pε(z)))2ψ(z)2dzdu,\displaystyle\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}=-\lim_{\varepsilon\to 0}\frac{4\pi}{\log\varepsilon}\int_{0}^{t}\int_{{\mathbb{R}}^{2}}\left(\mathscr{Z}^{\vartheta,\phi}_{u}(p_{\varepsilon}(\cdot-z))\right)^{2}\psi(z)^{2}\text{\rm d}z\text{\rm d}u, (1.11)

where

𝒵tϑ,ϕ(pε(x))=22ϕ(y)pε(zx)𝒵tϑ(dy,dz).\displaystyle\mathscr{Z}^{\vartheta,\phi}_{t}(p_{\varepsilon}(\cdot-x))=\int_{{\mathbb{R}}^{2}}\int_{{\mathbb{R}}^{2}}\phi(y)p_{\varepsilon}(z-x)\mathscr{Z}^{\vartheta}_{t}(\text{\rm d}y,\text{\rm d}z). (1.12)

and (1.11) is locally uniform convergence in probability.

Remark 1.12.

We remark that the martingale problem in Theorem 1.11 is ill-posed. Indeed, dropping the superscripts ϑ\vartheta in (1.10)-(1.12) does not change the martingale problem. However, we find that

𝒵tϑ,ϕ(1)=tϑ,ϕ(1)𝑑tϑ,ϕ(1)=𝒵tϑ,ϕ(1)\displaystyle\mathscr{Z}_{t}^{\vartheta,\phi}(1)=\mathscr{M}_{t}^{\vartheta,\phi}(1)\overset{d}{\not=}\mathscr{M}_{t}^{\vartheta^{\prime},\phi}(1)=\mathscr{Z}_{t}^{\vartheta^{\prime},\phi}(1)

for ϑϑ\vartheta\not=\vartheta^{\prime} by looking at their variances (see (2.13)). In particular, we know that the deterministic process

𝒵t,ϕ(ψ):=2dx2dyϕ(x)pt(x,y)ψ(y)\displaystyle\mathscr{Z}_{t}^{-\infty,\phi}(\psi):=\int_{{\mathbb{R}}^{2}}\text{\rm d}x\int_{{\mathbb{R}}^{2}}\text{\rm d}y\phi(x)p_{t}(x,y)\psi(y)

satisfies (1.11) with the constant quadratic variation. Thus, the martingale problem (1.10)-(1.12) has a family of solutions {𝒵ϑ,()}ϑ{}\{\mathscr{Z}^{\vartheta,\cdot}(\cdot)\}_{\vartheta\in\mathbb{R}\cup\{-\infty\}}.

Remark 1.13.

Let uu be a formal solution to the stochastic heat equation

tu=12Δu+λuW˙,u(0,x)=ϕ(x).\displaystyle\partial_{t}u=\frac{1}{2}\Delta u+\lambda u\dot{W},\quad u(0,x)=\phi(x).

and suppose it has a continuous density. Then, the quadratic variation process 2u(,x)ψ(x)dxt\langle\int_{{\mathbb{R}}^{2}}u(\cdot,x)\psi(x)\text{\rm d}x\rangle_{t} is formally given by

0t2λ2u2(s,x)ψ(x)2dxds\displaystyle\int_{0}^{t}\int_{{\mathbb{R}}^{2}}\lambda^{2}u^{2}(s,x)\psi(x)^{2}\text{\rm d}x\text{\rm d}s (1.13)

due to the effect of space-time white noise. Theorem 1.11 gives the rigorous definition of (1.13).

Actually, [15] shows that 𝒵tϑ,ϕ(dy)\mathscr{Z}^{\vartheta,\phi}_{t}(\text{\rm d}y) is not absolutely continuous with respect to Lebesgue measure, and hence 𝒵tϑ,ϕ(pε(x))\mathscr{Z}_{t}^{\vartheta,\phi}(p_{\varepsilon}(\cdot-x)) diverges at some points. So we need a renormalization factor 1log1ε\frac{1}{\log\frac{1}{\varepsilon}} in (1.11).

Remark 1.14.

For a usual super-Brownian motions {Xt}t0\{X_{t}\}_{t\geq 0}, their quadratic variation X(ψ)t\langle X(\psi)\rangle_{t} is given by

X(ψ)t=γ0tXs(ψ2)dst0\displaystyle\langle X(\psi)\rangle_{t}=\gamma\int_{0}^{t}X_{s}(\psi^{2})\text{\rm d}s\quad t\geq 0

for some γ>0\gamma>0 which is explicitly determined by the measure valued process {Xt}t0\{X_{t}\}_{t\geq 0}.

Just on the one hand, the quadratic variation for super-Brownian motion with a single point catalyst is represented by the density field which is given by the limit of 0tXs(pε(y))\int_{0}^{t}X_{s}(p_{\varepsilon}(\cdot-y)) for d=1d=1 [16].

By the definition of the cross variation of the martingales, we have the following.

Corollary 1.15.

Let ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}), ψ1,ψ2Cb2(2)\psi_{1},\psi_{2}\in C_{b}^{2}({\mathbb{R}}^{2}), and ϑ\vartheta\in{\mathbb{R}}. Let tϑ,ϕ(ψi)\mathscr{M}^{\vartheta,\phi}_{t}(\psi_{i}) (i=1,2i=1,2) be the martingales defined by (1.10) for ψ1,ψ2\psi_{1},\psi_{2}. Then, we have

ϑ,ϕ(ψ1),ϑ,ϕ(ψ2)t=limε04πlogε0t2(𝒵uϑ,ϕ(pε(z)))2ψ1(z)ψ2(z)dzdu,\displaystyle\left\langle\mathscr{M}^{\vartheta,\phi}(\psi_{1}),\mathscr{M}^{\vartheta,\phi}(\psi_{2})\right\rangle_{t}=-\lim_{\varepsilon\to 0}\frac{4\pi}{\log\varepsilon}\int_{0}^{t}\int_{{\mathbb{R}}^{2}}\left(\mathscr{Z}^{\vartheta,\phi}_{u}(p_{\varepsilon}(\cdot-z))\right)^{2}\psi_{1}(z)\psi_{2}(z)\text{\rm d}z\text{\rm d}u, (1.14)

where (1.14) is locally uniform convergence in probability.

Remark 1.16.

Theorem 1.11 gives the semimartingale representation of 𝒵ϑ,ϕ(ψ)\mathscr{Z}^{\vartheta,\phi}(\psi). It is natural to consider Itô’s formula to f(𝒵tϑ,ϕ(ψ))f\left(\mathscr{Z}_{t}^{\vartheta,\phi}(\psi)\right) for a function fCb2()f\in C_{b}^{2}({\mathbb{R}}). Then, we have

f(𝒵tϑ,ϕ(ψ))\displaystyle f\left(\mathscr{Z}_{t}^{\vartheta,\phi}(\psi)\right) =f(𝒵0ϑ,ϕ(ψ))+0tf(𝒵sϑ,ϕ(ψ))𝒵sϑ,ϕ(12Δψ)ds\displaystyle=f\left(\mathscr{Z}_{0}^{\vartheta,\phi}(\psi)\right)+\int_{0}^{t}f^{\prime}\left(\mathscr{Z}_{s}^{\vartheta,\phi}(\psi)\right)\mathscr{Z}_{s}^{\vartheta,\phi}\left(\frac{1}{2}\Delta\psi\right)\text{\rm d}s
+0tf(𝒵sϑ,ϕ(ψ))dsϑ,ϕ(ψ)\displaystyle+\int_{0}^{t}f^{\prime}\left(\mathscr{Z}_{s}^{\vartheta,\phi}(\psi)\right)\text{\rm d}\mathscr{M}_{s}^{\vartheta,\phi}\left(\psi\right)
+120tf′′(𝒵sϑ,ϕ(ψ))dϑ,ϕ(ψ)s.\displaystyle+\frac{1}{2}\int_{0}^{t}f^{\prime\prime}\left(\mathscr{Z}_{s}^{\vartheta,\phi}(\psi)\right)\text{\rm d}\left\langle\mathscr{M}^{\vartheta,\phi}\left(\psi\right)\right\rangle_{s}.

However, it is not obvious whether

0tf′′(𝒵sϑ,ϕ(ψ))dϑ,ϕ(ψ)s=limε04πlogε0t2f′′(𝒵sϑ,ϕ(ψ))(𝒵sϑ,ϕ(pε(z)))2ψ(z)2dzds\displaystyle\int_{0}^{t}f^{\prime\prime}\left(\mathscr{Z}_{s}^{\vartheta,\phi}(\psi)\right)\text{\rm d}\left\langle\mathscr{M}^{\vartheta,\phi}\left(\psi\right)\right\rangle_{s}=-\lim_{\varepsilon\to 0}\frac{4\pi}{\log\varepsilon}\int_{0}^{t}\int_{{\mathbb{R}}^{2}}f^{\prime\prime}\left(\mathscr{Z}_{s}^{\vartheta,\phi}(\psi)\right)\left(\mathscr{Z}_{s}^{\vartheta,\phi}(p_{\varepsilon}(\cdot-z))\right)^{2}\psi(z)^{2}\text{\rm d}z\text{\rm d}s (1.15)

in probability holds. The absolute continuity of ϑ,ϕ(ψ)s\left\langle\mathscr{M}^{\vartheta,\phi}\left(\psi\right)\right\rangle_{s} in time with respect to the Lebesgue measure is not clear. It remains open.

Remark 1.17.

To formulate (KPZ) via Cole-Hopf transformation, we may look at log𝒵tϑ,ϕ(pε(x))\log\mathscr{Z}_{t}^{\vartheta,\phi}\left(p_{\varepsilon}(\cdot-x)\right) instead of log𝒵tϑ,ϕ(x)\log\mathscr{Z}_{t}^{\vartheta,\phi}\left(x\right) since the latter is ill-posed. Since 𝒵ϑ,ϕ(pε(x))0\mathscr{Z}^{\vartheta,\phi}(p_{\varepsilon}(\cdot-x))\to 0 for Lebesgue almost everywhere x2x\in{\mathbb{R}}^{2} a.s. [15, (10.9)],

log𝒵ϑ,ϕ(pε(x))\displaystyle\log\mathscr{Z}^{\vartheta,\phi}(p_{\varepsilon}(\cdot-x))\to-\infty

for Lebesgue almost everywhere x2x\in{\mathbb{R}}^{2} a.s.. Thus, we also need to introduce another renormalization for this approach: Find aεa_{\varepsilon} and bε(x)b_{\varepsilon}(x) such that for each ψCc2(2)\psi\in C_{c}^{2}({\mathbb{R}}^{2})

aε2(log𝒵tϑ,ϕ(pε(x))bε(x))ψ(x)dxt(ϕ,ψ).\displaystyle a_{\varepsilon}\int_{{\mathbb{R}}^{2}}\left(\log\mathscr{Z}^{\vartheta,\phi}_{t}(p_{\varepsilon}(\cdot-x))-b_{\varepsilon}(x)\right)\psi(x)\text{\rm d}x\to\exists\mathfrak{H}_{t}(\phi,\psi).

1.2.1 Martingale measure

For fixed ϑ\vartheta\in{\mathbb{R}} and ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}), let c2\mathcal{M}_{c}^{2} be the set of continuous t𝒵ϑ,ϕ\mathcal{F}^{\mathscr{Z}^{\vartheta,\phi}}_{t}-martingales MM with M0=0M_{0}=0 and 𝔼[Mt2]<\mathbb{E}[M^{2}_{t}]<\infty for each t>0t>0.

Then, we found that ϑ,ϕ\mathscr{M}^{\vartheta,\phi} maps a function in Cb2(2)C_{b}^{2}({\mathbb{R}}^{2}) to a process tϑ,ϕ(ψ)c2\mathscr{M}^{\vartheta,\phi}_{t}(\psi)\in\mathcal{M}_{c}^{2} linearly.

In the following theorem, we will see the extension of the map ϑ,ϕ:Cb2(2)c2\mathscr{M}^{\vartheta,\phi}:C_{b}^{2}({\mathbb{R}}^{2})\to\mathcal{M}_{c}^{2}.

Theorem 1.18.

Let ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}), and ϑ\vartheta\in{\mathbb{R}}. Then, there exists a unique linear extension of ϑ,ϕ:Cb2(2)c2\mathscr{M}^{\vartheta,\phi}:C_{b}^{2}({\mathbb{R}}^{2})\to\mathcal{M}^{2}_{c} to ϑ,ϕ:b(2)c2\mathscr{M}^{\vartheta,\phi}:\mathcal{B}_{b}({\mathbb{R}}^{2})\to\mathcal{M}^{2}_{c} such that

ϑ,ϕ(ψ)t=limε04πlogε0t2(𝒵sϑ,ϕ(pε(z)))2ψ(z)2dzds\displaystyle\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}=\lim_{\varepsilon\to 0}\frac{4\pi}{-\log\varepsilon}\int_{0}^{t}\int_{{\mathbb{R}}^{2}}\left(\mathscr{Z}_{s}^{\vartheta,\phi}\left(p_{\varepsilon}(\cdot-z)\right)\right)^{2}\psi(z)^{2}\text{\rm d}z\text{\rm d}s (1.16)

for each t>0t>0, where b(2)\mathcal{B}_{b}({\mathbb{R}}^{2}) is the set of bounded Borel measurable functions and (1.16) is locally uniform convergence in probability.

For ψ(x):=1A(x)\psi(x):=1_{A}(x) (A(2)A\in\mathcal{B}(\mathbb{R}^{2})), we denote by tϑ,ϕ(A):=tϑ,ϕ(1A)\mathscr{M}^{\vartheta,\phi}_{t}(A):=\mathscr{M}^{\vartheta,\phi}_{t}(1_{A}).

Remark 1.19.

Theorem 1.18 and Corollary 1.15 imply that ϑ,ϕ()\mathcal{M}^{\vartheta,\phi}(\cdot) would be an orthogonal martingale measure in the sense of Walsh [45, Chapter 2]. Indeed, for each A(2)A\in\mathcal{B}({\mathbb{R}}^{2}), tϑ,ϕ(A)\mathcal{M}^{\vartheta,\phi}_{t}(A) is a continuous martingale and if A,B(2)A,B\in\mathcal{B}({\mathbb{R}}^{2}) with AB=A\cap B=\emptyset, then

ϑ,ϕ(A),ϑ,ϕ(B)t=0a.s. for each t>0.\displaystyle\left\langle\mathscr{M}^{\vartheta,\phi}(A),\mathscr{M}^{\vartheta,\phi}(B)\right\rangle_{t}=0\quad\text{a.s.~{}for each $t>0$.}

Moreover, we can define the stochastic integral with respect to this martingale measure in the general theory in [45]. However, we don’t discuss it in this paper.

1.3 Organization of the paper

In section 2, we review some results related to the analysis of moments of 𝖹\mathsf{Z} from [12, 15]. In section 3, we give an outline of the proof of Theorem 1.9 and Theorem 1.11. Section 4.1-6 are devoted to the detailed proofs. In section 7, we will discuss the regularity of the critical 2d2d stochastic heat flow.

2 Variance and its limit

In this section, we will look at the variances of 𝖹N;tβN,ϕ(ψ)\mathsf{Z}_{N;t}^{\beta_{N},\phi}(\psi).

First, it is easy to see that

𝔼[𝖹N;tβN,ϕ(ψ)]=1Nx2y2ϕ(xN)qNt(yx)ψ(yN)\displaystyle\mathbb{E}\left[\mathsf{Z}_{N;t}^{\beta_{N},\phi}(\psi)\right]=\frac{1}{N}\sum_{x\in{\mathbb{Z}}^{2}}\sum_{y\in{\mathbb{Z}}^{2}}\phi\left(\frac{x}{\sqrt{N}}\right)q_{Nt}(y-x)\psi\left(\frac{y}{\sqrt{N}}\right)

and

limN𝔼[𝖹N;tβN,ϕ(ψ)]2ϕ(x)dx2pt(yx)ψ(y)dy.\displaystyle\lim_{N\to\infty}\mathbb{E}\left[\mathsf{Z}_{N;t}^{\beta_{N},\phi}(\psi)\right]\to\int_{{\mathbb{R}}^{2}}\phi(x)\text{\rm d}x\int_{{\mathbb{R}}^{2}}p_{t}(y-x)\psi(y)\text{\rm d}y. (2.1)

Also, the standard L2L^{2}-moment method for DPRE yields

𝔼[𝖹N;tβN,ϕ(ψ)2]\displaystyle\mathbb{E}\left[\mathsf{Z}_{N;t}^{\beta_{N},\phi}(\psi)^{2}\right]
=1N2y,y2ψ(yN)ψ(yN)Ey,y[e(λ(2βN)2λ(βN))i=1Nt1{Si=Si}ψ(SNtN)ψ(SNtN)],\displaystyle=\frac{1}{N^{2}}\sum_{y,y^{\prime}\in{\mathbb{Z}}^{2}}\psi\left(\frac{y}{\sqrt{N}}\right)\psi\left(\frac{y^{\prime}}{\sqrt{N}}\right)E^{\otimes}_{y,y^{\prime}}\left[e^{(\lambda(2\beta_{N})-2\lambda(\beta_{N}))\sum_{i=1}^{\lfloor Nt\rfloor}1\{S_{i}=S_{i}^{\prime}\}}\psi\left(\frac{S_{Nt}}{\sqrt{N}}\right)\psi\left(\frac{S^{\prime}_{Nt}}{\sqrt{N}}\right)\right],

where (S,Px)(S,P_{x}) and (S,Px)(S^{\prime},P_{x^{\prime}}) are independent random walks starting at xx and xx^{\prime} whose increments has the same law with ξ\xi, respectively. Since

e(λ(2βN)2λ(βN))i=1Nt1{Si=Si}\displaystyle e^{(\lambda(2\beta_{N})-2\lambda(\beta_{N}))\sum_{i=1}^{\lfloor Nt\rfloor}1\{S_{i}=S_{i}^{\prime}\}} =i=1Nt(1+σN21{Si=Si})\displaystyle=\prod_{i=1}^{Nt}\left(1+\sigma_{N}^{2}1\{S_{i}=S_{i}^{\prime}\}\right)
=1+k=1NtσN2k1i1<<ikNtj=1k1{Sij=Sij},\displaystyle=1+\sum_{k=1}^{\lfloor Nt\rfloor}\sigma_{N}^{2k}\sum_{1\leq i_{1}<\dots<i_{k}\leq\lfloor Nt\rfloor}\prod_{j=1}^{k}1\{S_{i_{j}}=S_{i_{j}}^{\prime}\}, (2.2)

we have

Var(𝖹N;tβN,ϕ(ψ))\displaystyle\mathrm{Var}(\mathsf{Z}^{\beta_{N},\phi}_{N;t}(\psi))
=1N2x0,x02ϕ(x0N)ϕ(x0N)k=1NtσN2k0=i0<i1<<ikNtx1,,xk2\displaystyle=\frac{1}{N^{2}}\sum_{x_{0},x_{0}^{\prime}\in{\mathbb{Z}}^{2}}\phi\left(\frac{x_{0}}{\sqrt{N}}\right)\phi\left(\frac{x_{0}^{\prime}}{\sqrt{N}}\right)\sum_{k=1}^{Nt}\sigma_{N}^{2k}\sum_{0=i_{0}<i_{1}<\dots<i_{k}\leq\left\lfloor{Nt}\right\rfloor}\sum_{x_{1},\dots,x_{k}\in{\mathbb{Z}}^{2}}
qi0,i1(x0,x1)2j=2k(qij1,ij(xj1,xj)2)qik,Nt(xk,y)qik,Nt(xk,y)ψ(yN)ψ(yN),\displaystyle\qquad q_{i_{0},i_{1}}(x_{0},x_{1})^{2}\prod_{j=2}^{k}\left(q_{i_{j-1},i_{j}}(x_{j-1},x_{j})^{2}\right)q_{i_{k},\left\lfloor{Nt}\right\rfloor}(x_{k},y)q_{i_{k},\left\lfloor{Nt}\right\rfloor}(x_{k},y^{\prime})\psi\left(\frac{y}{\sqrt{N}}\right)\psi\left(\frac{y^{\prime}}{\sqrt{N}}\right),

where we set k=21=1\prod_{k=2}^{1}\cdots=1 and qi,j(x,y)=qji(yx)q_{i,j}(x,y)=q_{j-i}(y-x) for 0ij0\leq i\leq j and x,y2x,y\in{\mathbb{Z}}^{2}.

To see this quantity in detail, we use the following weighted local renewal functions introduced in [11] but we change the definition a little bit: For each N1N\geq 1,

UN(n,x)\displaystyle U_{N}(n,x) =σN4q0,n(0,x)2\displaystyle=\sigma_{N}^{4}q_{0,n}(0,x)^{2}
+k1(σN2)k+20<n1<<nk<nx1,,xk2q0,n1(0,x1)2(j=2kqnj1,nj(xj1,xj)2)qnk,n(xk,x)2n1\displaystyle\quad+\sum_{k\geq 1}(\sigma_{N}^{2})^{k+2}\sum_{\begin{smallmatrix}0<n_{1}<\dots<n_{k}<n\\ x_{1},\dots,x_{k}\in{\mathbb{Z}}^{2}\end{smallmatrix}}q_{0,n_{1}}(0,x_{1})^{2}\left(\prod_{j=2}^{k}q_{n_{j-1},n_{j}}(x_{j-1},x_{j})^{2}\right)q_{n_{k},n}(x_{k},x)^{2}\qquad n\geq 1 (2.3)
UN(0,x)\displaystyle U_{N}(0,x) =σN2δx,0=σN21{x=0}\displaystyle=\sigma_{N}^{2}\delta_{x,0}=\sigma_{N}^{2}1_{\{x=0\}} (2.4)
and
UN(n)\displaystyle U_{N}(n) =x2UN(n,x).\displaystyle=\sum_{x\in{\mathbb{Z}}^{2}}U_{N}(n,x). (2.5)

Then,

Var(𝖹N;tβN,ϕ(ψ))\displaystyle\mathrm{Var}(\mathsf{Z}^{\beta_{N},\phi}_{N;t}(\psi)) =1N2x0,x02ϕ(x0N)ϕ(x0N)qi(x0,x)qi(x0,x)\displaystyle=\frac{1}{N^{2}}\sum_{x_{0},x_{0}^{\prime}\in{\mathbb{Z}}^{2}}\phi\left(\frac{x_{0}}{\sqrt{N}}\right)\phi\left(\frac{x_{0}^{\prime}}{\sqrt{N}}\right)q_{i}(x_{0},x)q_{i}(x_{0}^{\prime},x)
1ijNtx,y,y,z2UN(ji,zx)qj,Nt(z,y)qj,Nt(z,y)ψ(yN)ψ(yN).\displaystyle\cdot\sum_{\begin{smallmatrix}1\leq i\leq j\leq\left\lfloor{Nt}\right\rfloor\\ x,y,y^{\prime},z\in{\mathbb{Z}}^{2}\end{smallmatrix}}U_{N}(j-i,z-x)q_{j,\left\lfloor{Nt}\right\rfloor}(z,y)q_{j,\left\lfloor{Nt}\right\rfloor}(z,y^{\prime})\psi\left(\frac{y}{\sqrt{N}}\right)\psi\left(\frac{y^{\prime}}{\sqrt{N}}\right). (2.6)

Thus, we can expect that UN(n,x)U_{N}(n,x) plays a key role in controlling the modulus of continuity of {𝖹N;tβN,ϕ(ψ)}t0\{\mathsf{Z}_{N;t}^{\beta_{N},\phi}(\psi)\}_{t\geq 0}.

We review some known results of UNU_{N} and 𝖹N;tβN,ϕ(ψ)\mathsf{Z}_{N;t}^{\beta_{N},\phi}(\psi) obtained in [11, 12]. We define

fs(t)={sts1eγsΓ(s+1)t(0,1]sts1eγsΓ(s+1)sts10t1fs(a)(1+a)2dat(1,)\displaystyle f_{s}(t)=\begin{cases}\displaystyle\frac{st^{s-1}e^{-\gamma s}}{\Gamma(s+1)}\qquad&t\in(0,1]\\ \displaystyle\frac{st^{s-1}e^{-\gamma s}}{\Gamma(s+1)}-st^{s-1}\int_{0}^{t-1}\frac{f_{s}(a)}{(1+a)^{2}}\text{\rm d}a\qquad&t\in(1,\infty)\end{cases}

for s(0,)s\in(0,\infty), and we set

Gϑ(t):=0eϑsfs(t)ds\displaystyle G_{\vartheta}(t):=\int_{0}^{\infty}e^{\vartheta s}f_{s}(t)\text{\rm d}s
Gϑ(t,x):=Gϑ(t)pt2(x),\displaystyle G_{\vartheta}(t,x):=G_{\vartheta}(t)p_{\frac{t}{2}}(x),

for t(0,)t\in(0,\infty) and x2x\in{\mathbb{R}}^{2}. In particular,

Gϑ(t)=0e(ϑγ)ssts1Γ(s+1)dsfor t(0,1].\displaystyle G_{\vartheta}(t)=\int_{0}^{\infty}\frac{e^{(\vartheta-\gamma)s}st^{s-1}}{\Gamma(s+1)}\text{\rm d}s\quad\text{for }t\in(0,1]. (2.7)
Theorem 2.1.

([11, Theorem 1.4, Theorem 2.3],[15, Proposition 8.4]) Suppose ϑ\vartheta\in{\mathbb{R}}. Then,

UN(n)=σN2logNNGϑ(nN)(1+o(1)),uniformly for δNnTN.\displaystyle U_{N}(n)=\frac{\sigma_{N}^{2}\log N}{N}G_{\vartheta}\left(\frac{n}{N}\right)\left(1+o(1)\right),\quad\textrm{uniformly for $\delta N\leq n\leq TN$.} (2.8)

for 0<δ<T<0<\delta<T<\infty. Also,

UN(n)CσN2logNNGϑ(nN),1nTN\displaystyle U_{N}(n)\leq C\frac{\sigma_{N}^{2}\log N}{N}G_{\vartheta}\left(\frac{n}{N}\right),\quad 1\leq n\leq TN (2.9)
and
UN(n,x)=σN2logNN2Gϑ(nN,xN)(1+o(1)),\displaystyle U_{N}(n,x)=\frac{\sigma_{N}^{2}\log N}{N^{2}}G_{\vartheta}\left(\frac{n}{N},\frac{x}{\sqrt{N}}\right)\left(1+o(1)\right),
uniformly for δNnN\delta N\leq n\leq N and |x|Nδ|x|\leq\frac{\sqrt{N}}{\delta}. (2.10)

In this paper, we often omit the subscript NN and denote by

Um,n(x,y)=UN(nm,yx),Um,n=y2Um,n(x,y)\displaystyle U_{m,n}(x,y)=U_{N}(n-m,y-x),U_{m,n}=\sum_{y\in{\mathbb{Z}}^{2}}U_{m,n}(x,y)

for 0mn<0\leq m\leq n<\infty and x,y2x,y\in{\mathbb{Z}}^{2}.

Proposition 2.2.

[11, Proposition 1.6], [12, Proposition 2.3], [15, Proposition 8.2] For fixed ϑ\vartheta\in{\mathbb{R}},

Gϑ(t)=1t(log1t)2+2ϑt(log1t)3+O(1t(1t)4),as t0.\displaystyle G_{\vartheta}(t)=\frac{1}{t\left(\log\frac{1}{t}\right)^{2}}+\frac{2\vartheta}{t\left(\log\frac{1}{t}\right)^{3}+O\left(\frac{1}{t\left(\frac{1}{t}\right)^{4}}\right)},\quad\text{as $t\to 0$}.

Also, for T>0T>0, there exists cϑ,T(0,)c_{\vartheta,T}\in(0,\infty) such that

Gϑ(t)G^ϑ,T(t):=cϑ,Tt(loge2Tt)2,t(0,T].\displaystyle G_{\vartheta}(t)\leq\widehat{G}_{\vartheta,T}(t):=\frac{c_{\vartheta,T}}{t\left(\log\frac{e^{2}T}{t}\right)^{2}},\quad t\in(0,T]. (2.11)

In particular, G^ϑ,T\widehat{G}_{\vartheta,T} is decreasing in t(0,T]t\in(0,T].

Combining (2.9) and (2.11), we obtain that

UN(n)1NCϑ,TnN(loge2TNn)2for 1nTN.\displaystyle U_{N}(n)\leq\frac{1}{N}\frac{C_{\vartheta,T}}{\frac{n}{N}\left(\log\frac{e^{2}TN}{n}\right)^{2}}\quad\text{for }1\leq n\leq TN. (2.12)

The following is a modification of Theorem 1.2 in [12] or Theorem 6.1 in [15].

Theorem 2.3.

Let ϑ\vartheta\in{\mathbb{R}}. Suppose βN\beta_{N} satisfies (1.8). Then, we have for each ϕCc(2)\phi\in C_{c}({\mathbb{R}}^{2}) and ψCb(2)\psi\in C_{b}({\mathbb{R}}^{2})

Var(𝒵tϑ,ϕ(ψ))=limNVar(𝖹N;tβN,ϕ(ψ))=4π0<u<v<tdudv2×2dxdyΦu(x)2Gϑ(vu,yx)Ψtv(y)2,\displaystyle\mathrm{Var}(\mathscr{Z}^{\vartheta,\phi}_{t}(\psi))=\lim_{N\to\infty}\mathrm{Var}(\mathsf{Z}_{N;t}^{\beta_{N},\phi}(\psi))=4\pi\int_{0<u<v<t}\text{\rm d}u\text{\rm d}v\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}x\text{\rm d}y\Phi_{u}(x)^{2}G_{\vartheta}(v-u,y-x)\Psi_{t-v}(y)^{2}, (2.13)

where we set Φs(x)=2ϕ(y)ps(xy)dy\Phi_{s}(x)=\int_{{\mathbb{R}}^{2}}\phi(y)p_{s}(x-y)\text{\rm d}y and Ψs(x)=2ψ(y)ps(xy)dy\Psi_{s}(x)=\int_{{\mathbb{R}}^{2}}\psi(y)p_{s}(x-y)\text{\rm d}y for s>0s>0 and x2x\in{\mathbb{R}}^{2}.

The higher moments of 𝒵tϑ,ϕ(ψ)\mathscr{Z}^{\vartheta,\phi}_{t}(\psi) are given explicitly in [25, Theorem 1.1] and [15, Theorem 9.6]. To write it, we prepare some notations: Let h2h\geq 2 be an integer. For I={i,j}I=\{i,j\} (1i<jh1\leq i<j\leq h), we wefine

(2)Ih:={(𝐱)=(x1,,xh)(2)h:xi=xj}\displaystyle({\mathbb{R}}^{2})^{h}_{I}:=\{(\mathbf{x})=(x_{1},\dots,x_{h})\in({\mathbb{R}}^{2})^{h}:x_{i}=x_{j}\}

which is identified with (2)h({\mathbb{R}}^{2})^{h}. We denote by ∫⋯∫(2)Ihf(𝐱)d𝐱I\idotsint_{({\mathbb{R}}^{2})^{h}_{I}}f(\mathbf{x})\text{\rm d}\mathbf{x}_{I} the integral of the integrable function on (2)Ih({\mathbb{R}}^{2})^{h}_{I} with respect to Lebesgue measure.

We define

Gϑ,tI(𝐱,𝐲):=({1,,h}\Ipt(yx))Gϑ(t,yixi)\displaystyle G^{I}_{\vartheta,t}(\mathbf{x},\mathbf{y}):=\left(\prod_{\ell\in\{1,\dots,h\}\backslash I}p_{t}(y_{\ell}-x_{\ell})\right)G_{\vartheta}(t,y_{i}-x_{i})

for 𝐱,𝐲(2)Ih,t>0\mathbf{x},\mathbf{y}\in({\mathbb{R}}^{2})^{h}_{I},t>0, and I={i,j}I=\{i,j\}. Also, we define

𝒬tI,J(𝐲,𝐱)=i=1hpt(xiyi)𝐲(2)Ih,𝐱(2)Jh,\displaystyle\mathscr{Q}^{I,J}_{t}(\mathbf{y},\mathbf{x})=\prod_{i=1}^{h}p_{t}(x_{i}-y_{i})\qquad\mathbf{y}\in({\mathbb{R}}^{2})^{h}_{I},\mathbf{x}\in({\mathbb{R}}^{2})^{h}_{J},
𝒬t,J(𝐲,𝐱)=i=1hpt(xiyi)𝐲(2)h,𝐱(2)Jh,\displaystyle\mathscr{Q}^{*,J}_{t}(\mathbf{y},\mathbf{x})=\prod_{i=1}^{h}p_{t}(x_{i}-y_{i})\qquad\mathbf{y}\in({\mathbb{R}}^{2})^{h},\mathbf{x}\in({\mathbb{R}}^{2})^{h}_{J},
𝒬tI,(𝐲,𝐱)=i=1hpt(xiyi)𝐲(2)Ih,𝐱(2)h\displaystyle\mathscr{Q}^{I,*}_{t}(\mathbf{y},\mathbf{x})=\prod_{i=1}^{h}p_{t}(x_{i}-y_{i})\qquad\mathbf{y}\in({\mathbb{R}}^{2})^{h}_{I},\mathbf{x}\in({\mathbb{R}}^{2})^{h}

for t>0t>0, and I,JI,J with |I|=|J|=2|I|=|J|=2.

Theorem 2.4.

Fix ϑ\vartheta\in{\mathbb{R}}. Let h2h\geq 2 be an integer. Then, for ϕ1,,ϕhCc(2)\phi_{1},\dots,\phi_{h}\in C_{c}({\mathbb{R}}^{2}), ψ1,,ψhCb(2)\psi_{1},\dots,\psi_{h}\in C_{b}({\mathbb{R}}^{2}), and t0t\geq 0

E[i=1h𝒵tϕi,ϑ(ψi)]=∫⋯∫(2)h×(2)hφh(𝐳)𝒦t(h)(𝐳,𝐰)ψh(𝐰)d𝐳d𝐰,\displaystyle E\left[\prod_{i=1}^{h}\mathscr{Z}^{\phi_{i},\vartheta}_{t}(\psi_{i})\right]=\idotsint\limits_{({\mathbb{R}}^{2})^{h}\times({\mathbb{R}}^{2})^{h}}\varphi^{\otimes h}(\mathbf{z})\mathscr{K}^{(h)}_{t}(\mathbf{z},\mathbf{w})\psi^{\otimes h}(\mathbf{w})\text{\rm d}\mathbf{z}\text{\rm d}\mathbf{w}, (2.14)

where we write 𝐱=(x1,,xh)(2)h\mathbf{x}=(x_{1},\dots,x_{h})\in({\mathbb{R}}^{2})^{h} and ϕh(𝐱):=i=1hϕ(xi)\phi^{\otimes h}(\mathbf{x}):=\prod_{i=1}^{h}\phi(x_{i}), and

𝒦t(h)(𝐳,𝐰):=\displaystyle\mathscr{K}_{t}^{(h)}(\mathbf{z},\mathbf{w}):=
1+m=1(4π)mI1,,Im{1,,h}|I|=2,II+1∫⋯∫0<a1<b1<<am<bm<t𝑑𝐚𝑑𝐛∫⋯∫𝐱(),𝐲()(2)Ihfor =1,,m=1md𝐱Id𝐲I\displaystyle 1+\sum_{m=1}^{\infty}(4\pi)^{m}\sum_{\begin{smallmatrix}I_{1},\dots,I_{m}\subset\{1,\dots,h\}\\ |I_{\ell}|=2,I_{\ell}\not=I_{\ell+1}\end{smallmatrix}}\quad\idotsint\limits_{0<a_{1}<b_{1}<\dots<a_{m}<b_{m}<t}d\mathbf{a}d\mathbf{b}\idotsint\limits_{\begin{smallmatrix}\mathbf{x}^{(\ell)},\mathbf{y}^{(\ell)}\in({\mathbb{R}}^{2})^{h}_{I_{\ell}}\\ \text{for }\ell=1,\dots,m\end{smallmatrix}}\prod_{\ell=1}^{m}\text{\rm d}\mathbf{x}_{I_{\ell}}\text{\rm d}\mathbf{y}_{I_{\ell}}
𝒬a1,I1(𝐳,𝐱(1))Gϑ,b1a1I1(𝐱(1),𝐲(1))(=2m𝒬ab1I1,I(𝐲(1),𝐱())Gϑ,baI(𝐱(),𝐲()))𝒬tbmIm,(𝐲(m),𝐰).\displaystyle\mathscr{Q}^{\asterisk,I_{1}}_{a_{1}}(\mathbf{z},\mathbf{x}^{(1)})G^{I_{1}}_{\vartheta,b_{1}-a_{1}}(\mathbf{x}^{(1)},\mathbf{y}^{(1)})\left(\prod_{\ell=2}^{m}\mathscr{Q}_{a_{\ell}-b_{\ell-1}}^{I_{\ell-1},I_{\ell}}(\mathbf{y}^{(\ell-1)},\mathbf{x}^{(\ell)})G^{I_{\ell}}_{\vartheta,b_{\ell}-a_{\ell}}(\mathbf{x}^{(\ell)},\mathbf{y}^{(\ell)})\right)\mathscr{Q}^{I_{m},\asterisk}_{t-b_{m}}(\mathbf{y}^{(m)},\mathbf{w}).

3 Proofs of Theorem 1.9 and Theorem 1.11

In this section, we will give the proofs of Theorem 1.9 and 1.11.

3.1 𝖹N;ϕ\mathsf{Z}_{N;\cdot}^{\phi} as measure valued process

We fix ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}).

Hereafter, we may assume that {ωn,x}n,x2\{\omega_{n,x}\}_{n\in{\mathbb{Z}},x\in{\mathbb{Z}}^{2}} i.i.d. random variables with

(ωn,x=1)=(ωn,x=1)=12.\displaystyle{\mathbb{P}}(\omega_{n,x}=1)={\mathbb{P}}(\omega_{n,x}=-1)=\frac{1}{2}. (3.1)

Indeed, the convergence to {𝒵tϑ(ϕ,ψ)}\{\mathscr{Z}_{t}^{\vartheta}(\phi,\psi)\} is independent of the choice of {ωn,x}\{\omega_{n,x}\} satisfying (1.1). In this case, it is easy to see that {ξn,xβ:=eβωn,xλ(β)1}(n,x)×2\left\{\xi_{n,x}^{\beta}:=e^{\beta\omega_{n,x}-\lambda(\beta)}-1\right\}_{(n,x)\in{\mathbb{N}}\times{\mathbb{Z}}^{2}} are i.i.d. random variables with (ξn,xβ=σβ)=(ξn,xβ=σβ)=12{\mathbb{P}}\left(\xi_{n,x}^{\beta}=\sigma_{\beta}\right)={\mathbb{P}}\left(\xi_{n,x}^{\beta}=-\sigma_{\beta}\right)=\frac{1}{2}, where we set σβ=tanh(β)\sigma_{\beta}=\tanh(\beta). In particular, we have

𝔼[(ξi,xβN)2m1]=0,𝔼[(ξi,xβN)2m]=σβNm\displaystyle\mathbb{E}\left[\left(\xi^{\beta_{N}}_{i,x}\right)^{2m-1}\right]=0,\quad\mathbb{E}\left[\left(\xi^{\beta_{N}}_{i,x}\right)^{2m}\right]=\sigma_{\beta_{N}}^{m} (3.2)

for mm\in{\mathbb{N}}.

(3.2) will help us estimating the chaos expansions of moments a bit (see Section 4), but it is not crucial.

We fix βN\beta_{N} as in (1.8). For simplicity of the notation, we write

𝖹tN,ϕ(dy)\displaystyle\mathsf{Z}^{N,\phi}_{t}(\text{\rm d}y) :=𝖹N;NtβN(ϕ,dy),𝖹tN,ϕ(ψ):=𝖹N;NtβN(ϕ,ψ).\displaystyle:=\mathsf{Z}^{\beta_{N}}_{N;\left\lfloor{Nt}\right\rfloor}(\phi,\text{\rm d}y),\quad\mathsf{Z}^{N,\phi}_{t}(\psi):=\mathsf{Z}_{N;\left\lfloor{Nt}\right\rfloor}^{\beta_{N}}(\phi,\psi).

Then, 𝖹tN,ϕ(dy)\mathsf{Z}^{N,\phi}_{t}(\text{\rm d}y) can be regarded as a measure-valued process with the initial value

𝖹0N,ϕ(dy)=1Ny2ϕ(yN)δyN(dy)\displaystyle\mathsf{Z}_{0}^{N,\phi}(\text{\rm d}y)=\frac{1}{N}\sum_{y\in{\mathbb{Z}}^{2}}\phi\left(\frac{y^{\prime}}{\sqrt{N}}\right)\delta_{\frac{y^{\prime}}{\sqrt{N}}}(\text{\rm d}y)
and
𝖹tN,ϕ(dy)\displaystyle\mathsf{Z}_{t}^{N,\phi}(\text{\rm d}y)
:=1Ny2E[ϕ(SNtN)exp(i=1Nt(βNωi,SNtiλ(βN)))|S0=y]δyN(dy).\displaystyle:=\frac{1}{N}\sum_{y^{\prime}\in{\mathbb{Z}}^{2}}E\left[\left.\phi\left(\frac{S_{\left\lfloor{Nt}\right\rfloor}}{\sqrt{N}}\right)\exp\left(\sum_{i=1}^{\left\lfloor{Nt}\right\rfloor}\left(\beta_{N}\omega_{i,S_{\left\lfloor{Nt}\right\rfloor-i}}-\lambda(\beta_{N})\right)\right)\right|S_{0}=y^{\prime}\right]\delta_{\frac{y}{\sqrt{N}}}(\text{\rm d}y^{\prime}).

We set

fN(x)=f(xN)\displaystyle f_{N}(x)=f\left(\frac{x}{\sqrt{N}}\right)
for fC(2)f\in C({\mathbb{R}}^{2}) and
Z¯N;nϕ(y)=E[ϕN(Sn)exp(i=1n1(βNωi,niλ(βN)))|S0=y].\displaystyle\overline{Z}^{\phi}_{N;n}(y)=E\left[\left.\phi_{N}\left(S_{n}\right)\exp\left(\sum_{i=1}^{n-1}\left(\beta_{N}\omega_{i,n-i}-\lambda(\beta_{N})\right)\right)\right|S_{0}=y\right].

Now, we look at {𝖹tN,ϕ(ϕ,ψ)}t0\{\mathsf{Z}_{t}^{N,\phi}(\phi,\psi)\}_{t\geq 0} as a discrete semimartingale as in the construction of super-Brownian motion from critical branching Brownian motion [41, II. 4].

Let {n}\{\mathcal{F}_{n}\} be a filtration generated by {ωi,x:x2,0in}\{\omega_{i,x}:x\in{\mathbb{Z}}^{2},0\leq i\leq n\} and we set ¯tN:=Nt\overline{\mathcal{F}}_{t}^{N}:=\mathcal{F}_{Nt} for t=kNt=\frac{k}{N} (k0k\in\mathbb{N}_{0}). Then, we have

𝖹k+1NN,ϕ(ψ)𝖹kNN,ϕ(ψ)\displaystyle\mathsf{Z}_{\frac{k+1}{N}}^{N,\phi}(\psi)-\mathsf{Z}_{\frac{k}{N}}^{N,\phi}(\psi)
=1Nx2ϕN(x)\displaystyle=\frac{1}{N}\sum_{x\in{\mathbb{Z}}^{2}}\phi_{N}(x)
×(y~2Ex[ek+1,Sk+1i=1kei,Si1Sk+1=y~]ψN(y~)y2Ex[i=1kei,Si1Sk=y]ψN(y))\displaystyle\times\left(\sum_{\tilde{y}\in{\mathbb{Z}}^{2}}E_{x}\left[e_{k+1,S_{k+1}}\prod_{i=1}^{k}e_{i,S_{i}}1_{S_{k+1}=\tilde{y}}\right]\psi_{N}(\tilde{y})-\sum_{y\in{\mathbb{Z}}^{2}}E_{x}\left[\prod_{i=1}^{k}e_{i,S_{i}}1_{S_{k}=y}\right]\psi_{N}(y)\right)
=1Nx2ϕN(x)y~2Ex[i=1kei,Si:Sk+1=y~]ψN(y~)(ek+1,y~1)\displaystyle=\frac{1}{N}\sum_{x\in{\mathbb{Z}}^{2}}\phi_{N}(x)\sum_{\tilde{y}\in{\mathbb{Z}}^{2}}E_{x}\left[\prod_{i=1}^{k}e_{i,S_{i}}:S_{k+1}=\tilde{y}\right]\psi_{N}(\tilde{y})\left(e_{k+1,\tilde{y}}-1\right)
+1Nx2ϕN(x)y2Z0,kβN(x,y)(y~2q1(y,y~)ψN(y~)ψN(y)).\displaystyle+\frac{1}{N}\sum_{x\in{\mathbb{Z}}^{2}}\phi_{N}(x)\sum_{y\in{\mathbb{Z}}^{2}}Z_{0,k}^{\beta_{N}}(x,y)\left(\sum_{\tilde{y}\in{\mathbb{Z}}^{2}}q_{1}(y,\tilde{y})\psi_{N}(\tilde{y})-\psi_{N}(y)\right).

where we set en,x=eβNωn,xλ(βN)=ξn,xβN+1e_{n,x}=e^{\beta_{N}\omega_{n,x}-\lambda(\beta_{N})}=\xi_{n,x}^{\beta_{N}}+1 for n0n\geq 0 and x2x\in{\mathbb{Z}}^{2}.

We define a discrete Laplacian ΔN\Delta_{N} by

ΔNψN(x)=N(y2q1(x,y)ψN(y)ψN(x))\displaystyle\Delta_{N}\psi_{N}\left(x\right)=N\left(\sum_{y\in{\mathbb{Z}}^{2}}q_{1}(x,y)\psi_{N}\left(y\right)-\psi_{N}\left(x\right)\right)

and we write

ΔMkN,ϕ(ψ)=1Ny~2Z¯N;k+1ϕ(y~)ψN(y~)(ek+1,y~1).\displaystyle\Delta M_{k}^{N,\phi}(\psi)=\frac{1}{N}\sum_{\tilde{y}\in{\mathbb{Z}}^{2}}\overline{Z}_{N;k+1}^{\phi}(\tilde{y})\psi_{N}(\tilde{y})\left(e_{k+1,\tilde{y}}-1\right).

Then, we have

𝖹k+1NN,ϕ(ψ)𝖹kNN,ϕ(ψ)=1N𝖹kNN,ϕ(ΔNψ)+ΔMkN,ϕ(ψ)\displaystyle\mathsf{Z}_{\frac{k+1}{N}}^{N,\phi}(\psi)-\mathsf{Z}_{\frac{k}{N}}^{N,\phi}(\psi)=\frac{1}{N}\mathsf{Z}_{\frac{k}{N}}^{N,\phi}\left(\Delta_{N}\psi\right)+\Delta M_{k}^{N,\phi}(\psi)

and hence

𝖹nNN,ϕ(ψ)\displaystyle\mathsf{Z}^{N,\phi}_{\frac{n}{N}}(\psi) =𝖹0N,ϕ(ψ)+1Nk=0n1𝖹kNN,ϕ(ΔNψ)+k=0n1ΔMkN,ϕ(ψ)\displaystyle=\mathsf{Z}_{0}^{N,\phi}(\psi)+\frac{1}{N}\sum_{k=0}^{n-1}\mathsf{Z}_{\frac{k}{N}}^{N,\phi}\left(\Delta_{N}\psi\right)+\sum_{k=0}^{n-1}\Delta M_{k}^{N,\phi}(\psi)
=𝖹0N,ϕ(ψ)+0nN𝖹tN,ϕ(ΔNψ)dt+k=0n1ΔMkN,ϕ(ψ).\displaystyle=\mathsf{Z}_{0}^{N,\phi}(\psi)+\int_{0}^{\frac{n}{N}}\mathsf{Z}_{t}^{N,\phi}\left(\Delta_{N}\psi\right)\text{\rm d}t+\sum_{k=0}^{n-1}\Delta M_{k}^{N,\phi}(\psi). (3.3)

We remark that

MnN,ϕ(ψ):=k=0n1ΔMkN,ϕ(ψ)\displaystyle M_{n}^{N,\phi}(\psi):=\sum_{k=0}^{n-1}\Delta M_{k}^{N,\phi}(\psi)

is an n\mathcal{F}_{n}-martingale with M0N,ϕ(ψ)=0M_{0}^{N,\phi}(\psi)=0 and the quadratic variation

MN,ϕ(ψ)n=σN2N2k=1ny2Z¯N;kϕ(y)2ψN(y)2.\displaystyle\left\langle M^{N,\phi}(\psi)\right\rangle_{n}=\frac{\sigma_{N}^{2}}{N^{2}}\sum_{k=1}^{n}\sum_{y\in{\mathbb{Z}}^{2}}\overline{Z}_{N;k}^{\phi}(y)^{2}\psi_{N}\left(y\right)^{2}.

In particular,

MN,ϕ(ψ)Nt\displaystyle\left\langle M^{N,\phi}(\psi)\right\rangle_{\left\lfloor{Nt}\right\rfloor} =σN2N2k=1Nty2Z¯N;kϕ(y)2ψN(y)2=:0NtNdsσN2Ny2Z¯N;Nsϕ(y)2ψN(y)2.\displaystyle=\frac{\sigma_{N}^{2}}{N^{2}}\sum_{k=1}^{\left\lfloor{Nt}\right\rfloor}\sum_{y\in{\mathbb{Z}}^{2}}{\overline{Z}}_{N;k}^{\phi}(y)^{2}\psi_{N}(y)^{2}=:\int_{0}^{\frac{\left\lfloor{Nt}\right\rfloor}{N}}\text{\rm d}s{\frac{\sigma_{N}^{2}}{N}\sum_{y\in{\mathbb{Z}}^{2}}{\overline{Z}}_{N;\left\lfloor{Ns}\right\rfloor}^{\phi}(y)^{2}\psi_{N}(y)^{2}}. (3.4)
Remark 3.1.

By the same argument as above, we can see that

𝔼[𝖹N,tβN(ϕ,ψ)|Nt1]=𝖹¯N,tβN(ϕ,ψ)\displaystyle\mathbb{E}\left[\left.\mathsf{Z}_{N,t}^{\beta_{N}}(\phi,\psi)\right|\mathcal{F}_{\left\lfloor{Nt}\right\rfloor-1}\right]=\overline{\mathsf{Z}}_{N,t}^{\beta_{N}}(\phi,\psi)

and hence

𝔼[(𝖹N,tβN(ϕ,ψ)𝖹¯N,tβN(ϕ,ψ))2]=𝔼[σN2N2y2Z¯N,kϕ(y)2ψN(y)2]0\displaystyle\mathbb{E}\left[\left(\mathsf{Z}_{N,t}^{\beta_{N}}(\phi,\psi)-\overline{\mathsf{Z}}_{N,t}^{\beta_{N}}(\phi,\psi)\right)^{2}\right]=\mathbb{E}\left[\frac{\sigma_{N}^{2}}{N^{2}}\sum_{y\in{\mathbb{Z}}^{2}}\overline{Z}_{N,k}^{\phi}(y)^{2}\psi_{N}(y)^{2}\right]\to 0

for each t0t\geq 0.

3.2 Continuity of 𝒵ϑ,ϕ\mathscr{Z}^{\vartheta,\phi}

We provide some important lemmas to prove Theorem 1.9 which are given in [41].

Definition 3.2.

Let EE be a Polish space. We say that the collection of processes {Xα:αI}\{X^{\alpha}:\alpha\in I\} with paths in D(E)D(E) is CC-relatively compact in D(E)D(E) if and only if it is relatively compact in D(E)D(E) and all weak limit points are a.s. continuous.

Definition 3.3.

Let EE be a Polish space. We say that DCb(E)D\subset C_{b}(E) is separating if and only if for any μ,νMF(E)\mu,\nu\in M_{F}(E), μ(ϕ)=ν(ϕ)\mu(\phi)=\nu(\phi) for all ϕD\phi\in D implies μ=ν\mu=\nu.

Then, Theorem 1.9 follows when we can verify the conditions (1) and (2) in the following theorem.

Theorem 3.4.

[41, Theorem II.4.1] Let EE be a Polish space and DCb(E)D\subset C_{b}(E) be a separating class in Cb(E)C_{b}(E) containing 11. A sequence of càdlàg MF(E)M_{F}(E)-valued processes {XN}\{X^{N}\} is CC-relatively compact in D(MF(E))D(M_{F}(E)) if and only if the following conditions hold:

  1. (1)

    For all ε>0\varepsilon>0, T>0T>0, there exists a compact set K=Kε,TK=K_{\varepsilon,T} in EE such that

    supNP(suptTXtN(Kε,Tc)>ε)<ε.\displaystyle\sup_{N}P\left(\sup_{t\leq T}X_{t}^{N}\left(K_{\varepsilon,T}^{c}\right)>\varepsilon\right)<\varepsilon.
  2. (2)

    For all ϕD\phi\in D, {XN(ϕ)}\left\{X^{N}(\phi)\right\} is CC-relatively compact in D()D({\mathbb{R}}).

If in addition, DD is closed under addition, then the above equivalence holds when ordinary relative compactness in DD replaces CC-relative compactness in both the hypothesis and conclusion.

Coming back to {𝖹tN,ϕ}\{\mathsf{Z}_{t}^{N,\phi}\}, we take D=Cb2(2)D=C_{b}^{2}({\mathbb{R}}^{2}) as the separating set in the proof of Theorem 1.9.

Proof of condition (1) in Theorem 3.4 for {𝖹tN,ϕ}\{\mathsf{Z}_{t}^{N,\phi}\}.

Fix ε>0\varepsilon>0 and T>0T>0. Then, by the invariance pinciple, we can take a compact set K2K\in{\mathbb{R}}^{2} such that

supN11Nx2ϕN(x)Px(SnNKc for some nNT)<ε2.\displaystyle\sup_{N\geq 1}\frac{1}{N}\sum_{x\in{\mathbb{Z}}^{2}}\phi_{N}\left(x\right)P_{x}\left(\frac{S_{n}}{\sqrt{N}}\in K^{c}\text{ for some $n\leq\left\lfloor{NT}\right\rfloor$}\right)<\varepsilon^{2}.

We regard 𝖹tN,ϕ\mathsf{Z}^{N,\phi}_{t} as a measure on the path space of random walk by

𝖹tN,ϕ(dS)=1Nx2ϕN(x)Ex[i=1Ntei,Si:dS].\displaystyle\mathsf{Z}^{N,\phi}_{t}(\text{\rm d}S)=\frac{1}{N}\sum_{x\in{\mathbb{Z}}^{2}}\phi_{N}(x)E_{x}\left[\prod_{i=1}^{\left\lfloor{Nt}\right\rfloor}e_{i,S_{i}}:\text{\rm d}S\right].

Then, it is clear that

𝖹tN,ϕ(Kc)𝖹tN,ϕ(SnNKc for some nNT)\displaystyle\mathsf{Z}^{N,\phi}_{t}(K^{c})\leq\mathsf{Z}^{N,\phi}_{t}\left(\frac{S_{n}}{\sqrt{N}}\in K^{c}\text{ for some $n\leq\left\lfloor{NT}\right\rfloor$}\right)

and the right-hand side is an n{\mathcal{F}}_{n}-martingale. Therefore, we have

(suptT𝖹tN,ϕ(Kc)>ε)\displaystyle{\mathbb{P}}\left(\sup_{t\leq T}\mathsf{Z}^{N,\phi}_{t}(K^{c})>\varepsilon\right) (suptT𝖹tN,ϕ(SnNKc for some nNT)>ε)\displaystyle\leq{\mathbb{P}}\left(\sup_{t\leq T}\mathsf{Z}^{N,\phi}_{t}\left(\frac{S_{n}}{\sqrt{N}}\in K^{c}\text{ for some $n\leq\left\lfloor{NT}\right\rfloor$}\right)>\varepsilon\right)
1ε𝔼[𝖹tN,ϕ(SnNKc for some nNT)]\displaystyle\leq\frac{1}{\varepsilon}\mathbb{E}\left[\mathsf{Z}^{N,\phi}_{t}\left(\frac{S_{n}}{\sqrt{N}}\in K^{c}\text{ for some $n\leq\left\lfloor{NT}\right\rfloor$}\right)\right]
=1ε1Nx2ϕN(x)Px(SnNKc for some nNT)<ε,\displaystyle=\frac{1}{\varepsilon}\frac{1}{N}\sum_{x\in{\mathbb{Z}}^{2}}\phi_{N}(x)P_{x}\left(\frac{S_{n}}{\sqrt{N}}\in K^{c}\text{ for some $n\leq NT$}\right)<\varepsilon, (3.5)

where we have used Doob’s maximal inequality in the second inequality. ∎

Next, we will verify (2) in Theorem 3.4 for {𝖹tN,ϕ}\{\mathsf{Z}_{t}^{N,\phi}\}. To see the CC-relative compactness in D([0,T],)D([0,T],{\mathbb{R}}) of {𝖹tN,ϕ(ψ)}t0\{\mathsf{Z}^{N,\phi}_{t}(\psi)\}_{t\geq 0}, it is enough to see the following two conditions

  1. (C-1)

    the CC-relative compactness of {0NtN𝖹sN,ϕ(ΔNψ)ds}\left\{\int_{0}^{\frac{\left\lfloor{Nt}\right\rfloor}{N}}\mathsf{Z}_{s}^{N,\phi}\left(\Delta_{N}\psi\right)\text{\rm d}s\right\} in D([0,T],)D([0,T],{\mathbb{R}}), and

  2. (C-2)

    the CC-relatively compactness of {MNtN,ϕ(ψ)}\left\{M^{N,\phi}_{\left\lfloor{Nt}\right\rfloor}(\psi)\right\} in D([0,T],)D([0,T],{\mathbb{R}}).

One can show (C-1) by the following lemma.

Lemma 3.5.

For any t0t\geq 0 and ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}),

supN1𝔼[suput𝖹uN,ϕ(1)2]<.\displaystyle\sup_{N\geq 1}\mathbb{E}\left[\sup_{u\leq t}\mathsf{Z}_{u}^{N,\phi}(1)^{2}\right]<\infty.
Proof of (C-1).

Fix T>0T>0. Then, we have for any ε>0\varepsilon>0 and δ>0\delta>0

(sup|ts|<δ0stT|NsNNtN𝖹uN,ϕ(ΔNψ)du|>ε)\displaystyle{\mathbb{P}}\left(\sup_{\begin{smallmatrix}|t-s|<\delta\\ 0\leq s\leq t\leq T\end{smallmatrix}}\left|\int_{\frac{\left\lfloor{Ns}\right\rfloor}{N}}^{\frac{\left\lfloor{Nt}\right\rfloor}{N}}\mathsf{Z}_{u}^{N,\phi}\left(\Delta_{N}\psi\right)\text{\rm d}u\right|>\varepsilon\right) (sup|ts|<δ0stT|NsNNtN𝖹uN,ϕ(1)du|ΔNψ>ε)\displaystyle\leq{\mathbb{P}}\left(\sup_{\begin{smallmatrix}|t-s|<\delta\\ 0\leq s\leq t\leq T\end{smallmatrix}}\left|\int_{\frac{\left\lfloor{Ns}\right\rfloor}{N}}^{\frac{\left\lfloor{Nt}\right\rfloor}{N}}\mathsf{Z}_{u}^{N,\phi}(1)\text{\rm d}u\right|\|\Delta_{N}\psi\|_{\infty}>\varepsilon\right)
(δsupuT𝖹uN,ϕ(1)ΔNψ>ε),\displaystyle\leq{\mathbb{P}}\left(\delta\sup_{u\leq T}\mathsf{Z}_{u}^{N,\phi}(1)\|\Delta_{N}\psi\|_{\infty}>\varepsilon\right), (3.6)

where we define f=supx2|f(x)|\|f\|_{\infty}=\sup_{x\in{\mathbb{R}}^{2}}|f(x)| for Cb(2)C_{b}({\mathbb{R}}^{2}). Thus, Lemma 3.5 implies that for any ε>0\varepsilon>0, there exists δ>0\delta>0 such that the right-hand side of (3.6) is smaller than ε\varepsilon. ∎

Proof of Lemma 3.5.

It is easy to see that for each N1N\geq 1, {𝖹kNN,ϕ(1)}k0\left\{\mathsf{Z}^{N,\phi}_{\frac{k}{N}}(1)\right\}_{k\geq 0} is an n{\mathcal{F}}_{n}-martingale so Doob’s maximal inequality yields that

𝔼[suput𝖹uN,ϕ(1)2]4𝔼[𝖹tN,ϕ(1)2]t0.\displaystyle\mathbb{E}\left[\sup_{u\leq t}\mathsf{Z}_{u}^{N,\phi}(1)^{2}\right]\leq 4\mathbb{E}\left[\mathsf{Z}_{t}^{N,\phi}(1)^{2}\right]\quad t\geq 0.

Thus, the proof is completed since we know from [12, Theorem 1.5], or from (2.1) and Theorem 2.3 that

𝔼[𝖹tN,ϕ(1)2](ϕ(x)dx)2+4π20<u<v<tΦu(x)2Gϑ(vu)dudvdx.\displaystyle\mathbb{E}\left[\mathsf{Z}_{t}^{N,\phi}(1)^{2}\right]\to\left(\int\phi(x)\text{\rm d}x\right)^{2}+4\pi\int_{{\mathbb{R}}^{2}}\int_{0<u<v<t}\Phi_{u}(x)^{2}G_{\vartheta}(v-u)\text{\rm d}u\text{\rm d}v\text{\rm d}x.

Thus, the proof of Theorem 1.9 has been completed once one can verify (C-2). To prove (C-2), we apply the general theory for CC-tightness of martingales given in [41, Lemma II.4.5.] or the conclusion of [30, Theorem VI.4.13, Theorem VI.3.26]

Lemma 3.6.

Let {MtN,tN}t=kN,k0\left\{M^{N}_{t},\mathcal{F}^{N}_{t}\right\}_{t=\frac{k}{N},k\in\mathbb{N}_{0}} be martingales with M0N=0M^{N}_{0}=0. Let

MNt:=0s<tE[(Ms+1NNMsN)2|sN],\displaystyle\left\langle M^{N}\right\rangle_{t}:=\sum_{0\leq s<t}E\left[\left(\left.M^{N}_{s+\frac{1}{N}}-M^{N}_{s}\right)^{2}\right|\mathcal{F}_{s}^{N}\right],

and extend MNM^{N}_{\cdot} and MN\langle M^{N}\rangle_{\cdot} to [0,)[0,\infty) as right-continuous step functions.

Then, the followings hold:

  1. (1)(1)

    Suppose the following two conditions:

    1. (C-2-i)(\mathrm{C}\text{-$2$-}\mathrm{i})

      {{MNt}t0:N1}\left\{\left\{\left\langle M^{N}_{\cdot}\right\rangle_{t}\right\}_{t\geq 0}:N\geq 1\right\} is CC-relatively compact in D()D({\mathbb{R}}).

    2. (C-2-ii)(\mathrm{C}\text{-$2$-}\mathrm{ii})
      sup0tT|Mt+1NNMtN|0in probability for all T1.\displaystyle\sup_{0\leq t\leq T}\left|M^{N}_{t+\frac{1}{N}}-M^{N}_{t}\right|\to 0\quad\text{in probability for all $T\geq 1$}. (3.7)

    Then. MNM^{N}_{\cdot} is CC-relatively compact in D()D({\mathbb{R}}).

  2. (2)(2)

    If, in addition,

    {(MsN)2+MNs2:N1}is uniformly integrable for all s[0,T],\displaystyle\left\{\left(M^{N}_{s}\right)^{2}+\left\langle M^{N}\right\rangle_{s}^{2}:N\geq 1\right\}\quad\text{is uniformly integrable for all $s\in[0,T]$}, (3.8)

    then MNkM^{N_{k}}_{\cdot}\Rightarrow\mathscr{M}_{\cdot} implies that \mathscr{M}_{\cdot} is a continuous L2L^{2}-martingale with respect to the filtration t\mathcal{F}_{t}^{\mathscr{M}} and that

    (MNk,MNk)(,)\displaystyle\left(M^{N_{k}}_{\cdot},\left\langle M^{N_{k}}\right\rangle_{\cdot}\right)\Rightarrow\left(\mathscr{M}_{\cdot},\left\langle\mathscr{M}\right\rangle_{\cdot}\right)

Once we verify (1)(1)(C-2-i)(\mathrm{C}\text{-$2$-}\mathrm{i}) and (1)(1)(C-2-ii)(\mathrm{C}\text{-$2$-}\mathrm{ii}) in Lemma 3.6 for our martingales MN,ϕ(ψ)M_{\cdot}^{N,\phi}(\psi), (C-2) follows.

3.2.1 Proof of (1)(1)(C-2-i)(\mathrm{C}\text{-$2$-}\mathrm{i}) for MN,ϕ(ψ)M^{N,\phi}(\psi)

To prove (1)(1)(C-2-i)(\mathrm{C}\text{-$2$-}\mathrm{i}) in Lemma 3.6 for MN,ϕM^{N,\phi}, we adapt the standard method.

Lemma 3.7.

For each T0T\geq 0, p>1p>1, ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}), and ψCb(2)\psi\in C_{b}({\mathbb{R}}^{2}), there exists C>0C>0 such that

𝔼[(MN,ϕ(ψ)NtMN,ϕ(ψ)Ns)2]C|ts|321p0stT.\displaystyle\mathbb{E}\left[\left(\left\langle M^{N,\phi}(\psi)\right\rangle_{\left\lfloor{Nt}\right\rfloor}-\left\langle M^{N,\phi}(\psi)\right\rangle_{\left\lfloor{Ns}\right\rfloor}\right)^{2}\right]\leq C|t-s|^{\frac{3}{2}-\frac{1}{p}}\quad 0\leq s\leq t\leq T. (3.9)

Then, applying the Garsia-Rodemich-Rumsey inequality [22, Theorem A.1], (1)(1)(C-2-i)(\mathrm{C}\text{-$2$-}\mathrm{i}) follows.

Remark 3.8.

Lemma 3.7 gives an upper bound |ts|32o(1)|t-s|^{\frac{3}{2}-o(1)}, and it is not probably optimal. Actually, we can see from [44, Lemma 4.3] that it is of order |ts|2o(1)|t-s|^{2-o(1)} for the continuous setting.

The proof of Lemma 3.7 will be given in Section 5.

Here is an easy estimate of the difference between the values of MN,ϕ(ψ)Nt\langle M^{N,\phi}(\psi)\rangle_{Nt}. For 0<s<t<T0<s<t<T,

14πlimNE[MN,ϕ(ψ)NtMN,ϕ(ψ)Ns]\displaystyle\frac{1}{4\pi}\lim_{N\to\infty}E[\langle M^{N,\phi}(\psi)\rangle_{Nt}-\langle M^{N,\phi}(\psi)\rangle_{Ns}]
20<u<v<tΦu(x)2Gϑ(vu)dudvdx20<u<v<sΦu(x)2Gϑ(vu)dudvdx\displaystyle\leq\int_{{\mathbb{R}}^{2}}\int_{0<u<v<t}\Phi_{u}(x)^{2}G_{\vartheta}(v-u)\text{\rm d}u\text{\rm d}v\text{\rm d}x-\int_{{\mathbb{R}}^{2}}\int_{0<u<v<s}\Phi_{u}(x)^{2}G_{\vartheta}(v-u)\text{\rm d}u\text{\rm d}v\text{\rm d}x
20<u<s<v<tΦu(x)2G^ϑ,T(vu)dudvdx+2s<u<v<tΦu(x)2G^ϑ,T(vu)dudvdx\displaystyle\leq\int_{{\mathbb{R}}^{2}}\int_{0<u<s<v<t}\Phi_{u}(x)^{2}\widehat{G}_{\vartheta,T}(v-u)\text{\rm d}u\text{\rm d}v\text{\rm d}x+\int_{{\mathbb{R}}^{2}}\int_{s<u<v<t}\Phi_{u}(x)^{2}\widehat{G}_{\vartheta,T}(v-u)\text{\rm d}u\text{\rm d}v\text{\rm d}x
Cϕ,ϑ,T,1tsloge2Ts+Cϕ,ϑ,T,2tsloge2Tts,\displaystyle\leq C_{\phi,\vartheta,T,1}\frac{t-s}{\log\frac{e^{2}T}{s}}+C_{\phi,\vartheta,T,2}\frac{t-s}{\log\frac{e^{2}T}{t-s}}, (3.10)

where we have used G^ϑ,T(vu)G^ϑ,T(su)\widehat{G}_{\vartheta,T}(v-u)\leq\widehat{G}_{\vartheta,T}(s-u) for 0<u<s<v<T0<u<s<v<T in the first term, and G^ϑ,T(t)=ddtcϑ,Tlog(2Tt)\widehat{G}_{\vartheta,T}(t)=\frac{\text{\rm d}}{\text{\rm d}t}\frac{c_{\vartheta,T}}{\log\left(\frac{2T}{t}\right)} for t(0,T]t\in(0,T] and

st1log(e2Ttu)dutse2T1log(e2Tts)as ts0\displaystyle\int_{s}^{t}\frac{1}{\log\left(\frac{e^{2}T}{t-u}\right)}\text{\rm d}u\sim\frac{t-s}{e^{2}T}\frac{1}{\log\left(\frac{e^{2}T}{t-s}\right)}\quad\text{as }t-s\searrow 0

in the second term.

3.2.2 Proof of (1)(1)(C-2-ii)(\mathrm{C}\text{-$2$-}\mathrm{ii}) for MN,ϕ(ψ)M^{N,\phi}(\psi)

To prove (1)(1)(C-2-ii)(\mathrm{C}\text{-$2$-}\mathrm{ii}) for MN,ϕ(ψ)M^{N,\phi}_{\cdot}(\psi), we first remark that

sup0kNt1|ΔMN,kϕ(ψ)|3k=0Nt1|ΔMN,kϕ(ψ)|3\displaystyle\sup_{0\leq k\leq\left\lfloor{Nt}\right\rfloor-1}\left|\Delta M_{N,k}^{\phi}(\psi)\right|^{3}\leq\sum_{k=0}^{\left\lfloor{Nt}\right\rfloor-1}\left|\Delta M_{N,k}^{\phi}(\psi)\right|^{3} (3.11)

To prove (3.7), we use the following Burkholder type inequality [41, (PSF) in p.152] and [4, Theorem 21.1].

Lemma 3.9.

Let f:[0,)[0,)f:[0,\infty)\to[0,\infty) be a continuous increasing function with f(0)=0f(0)=0 such that there exists c0(0,)c_{0}\in(0,\infty) such that f(2λ)c0f(λ)f(2\lambda)\leq c_{0}f(\lambda) for any λ0\lambda\geq 0.

Let {Mn}n0\{M_{n}\}_{n\geq 0} be an n{\mathcal{F}}_{n}-martingale. We set Mn=supkn|Mk|M_{n}^{*}=\sup_{k\leq n}|M_{k}|,

Mn:=k=1nE[(MkMk1)2|k1]+E[M02]\displaystyle\langle M\rangle_{n}:=\sum_{k=1}^{n}E\left[(M_{k}-M_{k-1})^{2}|{\mathcal{F}}_{k-1}\right]+E[M_{0}^{2}]
dn:=max1kn|MkMk1|.\displaystyle d_{n}^{*}:=\max_{1\leq k\leq n}|M_{k}-M_{k-1}|.

Then, we have

E[f(Mn)]c(E[f(Mn12)]+E[f(dn)]).\displaystyle E\left[f(M_{n}^{*})\right]\leq c\left(E\left[f\left(\langle M\rangle_{n}^{\frac{1}{2}}\right)\right]+E\left[f(d_{n}^{*})\right]\right).
Proof of (3.7) for MnN,ϕ(ψ)M_{n}^{N,\phi}(\psi).

Conditioned on n1{\mathcal{F}}_{n-1}, ΔMN,nϕ(ψ)\Delta M_{N,n}^{\phi}(\psi) is a sum of mean 0 independent random variables y2M~y(N,n)(ϕ,ψ)\sum_{y\in{\mathbb{Z}}^{2}}\widetilde{M}^{(N,n)}_{y}(\phi,\psi), where

M~y(N,n)(ϕ,ψ):=1NZ¯N,nNϕ(y)ψN(y)ξn,yβN.\displaystyle\widetilde{M}^{(N,n)}_{y}(\phi,\psi):=\frac{1}{N}\overline{Z}_{N,\frac{n}{N}}^{\phi}(y)\psi_{N}(y)\xi^{\beta_{N}}_{n,y}.

Let Λn\Lambda_{n} be the subset of 2{\mathbb{Z}}^{2} such that ΛnΛn+1\Lambda_{n}\subset\Lambda_{n+1}, |Λn|=n|\Lambda_{n}|=n for n0n\geq 0 and n1Λn=2\bigcup_{n\geq 1}\Lambda_{n}={\mathbb{Z}}^{2}. We define the filtration n,Λk=n1σ[ωn,x,xΛk]\mathcal{F}_{n,\Lambda_{k}}=\mathcal{F}_{n-1}\vee\sigma[\omega_{n,x},x\in\Lambda_{k}]. Then,

M~0(N,n)(ϕ,ψ):=0,M~k(N,n)(ϕ,ψ):=yΛkM~y(N,n)(ϕ,ψ)(k1)\displaystyle\widetilde{M}_{0}^{(N,n)}(\phi,\psi):=0,\quad\widetilde{M}_{k}^{(N,n)}(\phi,\psi):=\sum_{y\in\Lambda_{k}}\widetilde{M}_{y}^{(N,n)}(\phi,\psi)\quad(k\geq 1)

is n,Λk\mathcal{F}_{n,\Lambda_{k}}-martingale, and limkM~k(N,n)(ϕ,ψ)=ΔMnN,ϕ(ψ)\lim_{k\to\infty}\widetilde{M}_{k}^{(N,n)}(\phi,\psi)=\Delta M_{n}^{N,\phi}(\psi) since the summation is finite. Thus, we can apply Lemma 3.9 with f(λ)=λ3f(\lambda)=\lambda^{3} (c(f)=23c(f)=2^{3}) to {M~y(N,n)(ϕ,ψ)}k0\{\widetilde{M}^{(N,n)}_{y}(\phi,\psi)\}_{k\geq 0}. Also, we can see that

M~(N,n)(ϕ,ψ)k=yΛkσN2N2Z¯N,kϕ(y)2,\displaystyle\left\langle\widetilde{M}^{(N,n)}(\phi,\psi)\right\rangle_{k}=\sum_{y\in\Lambda_{k}}\frac{\sigma_{N}^{2}}{N^{2}}\overline{Z}_{N,k}^{\phi}(y)^{2}, |M~y(N,n)(ϕ,ψ)|=σNNZ¯N,nNϕ(y)|ψ(y)|\displaystyle\left|\widetilde{M}_{y}^{(N,n)}(\phi,\psi)\right|=\frac{\sigma_{N}}{N}\overline{Z}_{N,\frac{n}{N}}^{\phi}(y)|\psi(y)|

by (3.1). Thus, we obtain

E[|ΔMN,nϕ(ψ)|3]\displaystyle E\left[\left|\Delta M_{N,n}^{\phi}(\psi)\right|^{3}\right] c(f)E[(σN2N2y2Z¯N;nNϕ,(y)2ψ(yN)2)32]\displaystyle\leq c(f)E\left[\left(\frac{\sigma_{N}^{2}}{N^{2}}\sum_{y\in{\mathbb{Z}}^{2}}{\overline{Z}}_{N;\frac{n}{N}}^{\phi,}(y)^{2}\psi\left(\frac{y}{\sqrt{N}}\right)^{2}\right)^{\frac{3}{2}}\right]
+c(f)E[y2(σN2N2Z¯N;nNϕ(y)2ψ(yN)2)32]\displaystyle+c(f)E\left[\sum_{y\in{\mathbb{Z}}^{2}}\left(\frac{\sigma_{N}^{2}}{N^{2}}{\overline{Z}}_{N;\frac{n}{N}}^{\phi}(y)^{2}\psi\left(\frac{y}{\sqrt{N}}\right)^{2}\right)^{\frac{3}{2}}\right]
c(f)E[y22σN3N3Z¯N;nNϕ(y)3]\displaystyle\leq c(f)E\left[\sum_{y\in{\mathbb{Z}}^{2}}\frac{2\sigma_{N}^{3}}{N^{3}}{\overline{Z}}_{N;\frac{n}{N}}^{\phi}(y)^{3}\right]

where we have used (dn)31kndk3(d_{n}^{*})^{3}\leq\sum_{1\leq k\leq n}d_{k}^{3} in the first inequality.

We can see from [12, Lemma 6.1 (6.4)] that for any ε>0\varepsilon>0 and t0t\geq 0, there exists C>0C>0 such that

1nNty2E[Z¯N;nNϕ(y)3]CN52+ε.\displaystyle\sum_{1\leq n\leq Nt}\sum_{y\in{\mathbb{Z}}^{2}}E\left[{\overline{Z}}_{N;\frac{n}{N}}^{\phi}(y)^{3}\right]\leq CN^{\frac{5}{2}+\varepsilon}. (3.12)

Thus, combining this with (3.11), (3.7) follows.

Remark 3.10.

For (3.12), we see that

E[Z¯N;nNϕ(y)3]\displaystyle E\left[{\overline{Z}}_{N;\frac{n}{N}}^{\phi}(y)^{3}\right] =E[(Z¯N;nϕ(y)E[Z¯N;nNϕ(y)]+E[Z¯N;nNϕ(y)])3]\displaystyle=E\left[\left({\overline{Z}}_{N;n}^{\phi}(y)-E\left[{\overline{Z}}_{N;\frac{n}{N}}^{\phi}(y)\right]+E\left[{\overline{Z}}_{N;\frac{n}{N}}^{\phi}(y)\right]\right)^{3}\right]
=E[(Z¯N;nϕ(y)E[Z¯N;nNϕ(y)])3]+3E[(Z¯N;nNϕ(y)E[Z¯N;nNϕ(y)])2]E[Z¯N;nNϕ(y)]\displaystyle=E\left[\left({\overline{Z}}_{N;n}^{\phi}(y)-E\left[{\overline{Z}}_{N;\frac{n}{N}}^{\phi}(y)\right]\right)^{3}\right]+3E\left[\left({\overline{Z}}_{N;\frac{n}{N}}^{\phi}(y)-E\left[{\overline{Z}}_{N;\frac{n}{N}}^{\phi}(y)\right]\right)^{2}\right]E\left[{\overline{Z}}_{N;\frac{n}{N}}^{\phi}(y)\right]
+E[Z¯N;nNϕ(y)]3.\displaystyle+E\left[{\overline{Z}}_{N;\frac{n}{N}}^{\phi}(y)\right]^{3}.

Our setting of {ωn,x}\{\omega_{n,x}\}, in particular (3.2), allows us to use [12, Lemma 6.1 (6.4)] for an upper bound of the first term. More precisely, the expectation was divided into two terms, non-triple intersections and triple intersections, and the latter one vanishes under (3.2). The second term can be easily estimated by (2.6) and Lemma 2.1.

Thus, (1)(1) in Lemma 3.6 follows when we verify Lemma 3.7.

3.2.3 Proof of (2)(2) for MN,ϕ(ψ)M^{N,\phi}_{\cdot}(\psi)

Proof of (2)(2) in Lemma 3.6.

We use Lemma 3.9 again. Taking f(λ)=λ3f(\lambda)=\lambda^{3},

E[sup0sT|MN;Nsϕ(ψ)|3]c(f)E[MN;ϕ(ψ)NT32]+c(f)E[sup1kNT1|ΔMN;kϕ(ψ)|3].\displaystyle E\left[\sup_{0\leq s\leq T}\left|M_{N;Ns}^{\phi}(\psi)\right|^{3}\right]\leq c(f)E\left[\left\langle M_{N;\cdot}^{\phi}(\psi)\right\rangle_{NT}^{\frac{3}{2}}\right]+c(f)E\left[\sup_{1\leq k\leq NT-1}\left|\Delta M_{N;k}^{\phi}(\psi)\right|^{3}\right].

The second term is already estimated in the proof of (1)(1)(C-2-ii)(\mathrm{C}\text{-$2$-}\mathrm{ii}). Also, the expectation in the first term is dominated by

c(f)ψ(x)3E[(1kNT|ΔMN;kϕ(1)|)32]\displaystyle c(f)\|\psi(x)\|_{\infty}^{3}E\left[\left(\sum_{1\leq k\leq NT}\left|\Delta M_{N;k}^{\phi}(1)\right|\right)^{\frac{3}{2}}\right]

from [4, Lemma 16.1]. Also, combining Theorem 15.1 in [4] and Doob’s LpL^{p}-inequality, this is dominated by

CE[|MN;NTψ(1)|3]CE[𝖹N;T|ϕ|(1)3]+C𝖹N;0|ϕ|(1)3,\displaystyle CE\left[\left|M_{N;NT}^{\psi}(1)\right|^{3}\right]\leq CE\left[\mathsf{Z}_{N;T}^{|\phi|}(1)^{3}\right]+C\mathsf{Z}_{N;0}^{|\phi|}(1)^{3},

where we remark that if ψ1\psi\equiv 1, then MN,Nsϕ(1)=𝖹N;sϕ(1)𝖹N;0ϕ(1)M_{N,Ns}^{\phi}(1)=\mathsf{Z}_{N;s}^{\phi}(1)-\mathsf{Z}_{N;0}^{\phi}(1). We know that the right-hand side is bounded from [12, Theorem 1.4]. ∎

3.3 Martingale tϑ,ϕ(ψ)\mathscr{M}_{t}^{\vartheta,\phi}(\psi)

Suppose that {MNtN,ϕ(ψ)}\{M^{N,\phi}_{Nt}(\psi)\} satisfy all conditions in Lemma 3.6 so Theorem 1.9 follows. Theorem 1.4 implies that 𝖹N,ϕ𝒵ϑ,ϕ\mathsf{Z}^{N,\phi}\Rightarrow\mathscr{Z}^{\vartheta,\phi} in D(MF(2))D(M_{F}({\mathbb{R}}^{2})).

By the Skorokhod representation theorem, we may assume that 𝖹Nk,ϕ\mathsf{Z}^{N_{k},\phi} and 𝒵ϑ,ϕ\mathscr{Z}^{\vartheta,\phi} are defined on a common probability space and 𝖹Nk,ϕ𝒵ϑ,ϕ\mathsf{Z}^{N_{k},\phi}_{\cdot}\to\mathscr{Z}^{\vartheta,\phi} in D(MF(2))D(M_{F}({\mathbb{R}}^{2})) a.s.

Then, all terms in the righthand side of (3.3) (and hence {MNN,ϕ(ψ)}\displaystyle\left\{M^{N,\phi}_{N\cdot}(\psi)\right\}) converge almost surely. Indeed, 𝖹0N,ϕ(ψ)ϕ(x)ψ(x)dx\mathsf{Z}^{N,\phi}_{0}(\psi)\to\int\phi(x)\psi(x)\text{\rm d}x, and Taylor’s theorem implies that for each x2,yΣx\in{\mathbb{Z}}^{2},y\in\Sigma and NN\in{\mathbb{N}}, there exists c=cN,x,yc=c_{N,x,y} such that

ψN(x+y)ψN(x)=ψN(x)yN+12N(y1,y2)Hess(ψ)(x+cyN)(y1y2),\displaystyle\psi_{N}(x+y)-\psi_{N}(x)=\psi^{\prime}_{N}(x)\frac{y}{\sqrt{N}}+\frac{1}{2N}(y_{1},y_{2})\mathrm{Hess}(\psi)\left(\frac{x+cy}{\sqrt{N}}\right)\left(\begin{matrix}y_{1}\\ y_{2}\end{matrix}\right),

where Hess(ψ)\mathrm{Hess}(\psi) is the Hesse matrix of ψ\psi.

Thus, we have

yΣq1(y)(ψN(x+y)ψN(x))=12Δψ(x)+o(1)\displaystyle\sum_{y\in\Sigma}q_{1}(y)\left(\psi_{N}(x+y)-\psi_{N}(x)\right)=\frac{1}{2}\Delta\psi(x)+o(1)

uniformly in any compact set K2K\subset{\mathbb{R}}^{2}. Hence, we can see from (3.5) that

0NtN𝖹sNk,ϕ(ΔNϕ)ds120t𝒵sϑ,ϕ(Δψ)ds,a.s.\displaystyle\int_{0}^{\frac{\left\lfloor{Nt}\right\rfloor}{N}}\mathsf{Z}_{s}^{N_{k},\phi}(\Delta_{N}\phi)\text{\rm d}s\to\frac{1}{2}\int_{0}^{t}\mathscr{Z}_{s}^{\vartheta,\phi}(\Delta\psi)\text{\rm d}s,\quad\textrm{a.s.}

Thus, we found that MNN,ϕ(ψ)ϑ,ϕ(ψ)M^{N,\phi}_{N\cdot}(\psi)\Rightarrow\mathscr{M}^{\vartheta,\phi}_{\cdot}(\psi) in D([0,),)D([0,\infty),{\mathbb{R}}) and hence, Lemma 3.6 implies that

(MNN,ϕ(ψ),MN,ϕ(ψ)N)(ϑ,ϕ(ψ),ϑ,ϕ(ψ)),\displaystyle\left(M^{N,\phi}_{N\cdot}(\psi),\left\langle M^{N,\phi}(\psi)\right\rangle_{N\cdot}\right)\Rightarrow\left(\mathscr{M}^{\vartheta,\phi}(\psi),\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{\cdot}\right), (3.13)

and that ϑ,ϕ(ψ)\mathscr{M}^{\vartheta,\phi}_{\cdot}(\psi) is a continuous tϑ,ϕ(ψ)\mathcal{F}^{\mathscr{M}^{\vartheta,\phi}(\psi)}_{t}-martingale (not tϑ,ϕ(ψ)\mathcal{F}^{\mathscr{M}^{\vartheta,\phi}(\psi)}_{t}-martingale).

Therefore, the proof of Theorem 1.11 is completed when we proved the following two lemmas.

Lemma 3.11.

For any ϑ\vartheta\in{\mathbb{R}}, ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}), and ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}), tϑ,ϕ(ψ)\mathscr{M}^{\vartheta,\phi}_{t}(\psi) is a continuous t𝒵ϑ,ϕ\mathcal{F}_{t}^{\mathscr{Z}^{\vartheta,\phi}}-martingale.

Lemma 3.12.

For any ϑ\vartheta\in{\mathbb{R}}, ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}), and ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}),

θ,ϕ(ψ)t=limε04πlogε0t2𝒵sϑ,ϕ(pε(z))2ψ(z)2dzdsuniformly on [0,T] in probability\displaystyle\displaystyle\left\langle\mathscr{M}^{\theta,\phi}(\psi)\right\rangle_{t}=\lim_{\varepsilon\to 0}\frac{4\pi}{-\log\varepsilon}\int_{0}^{t}\int_{{\mathbb{R}}^{2}}\mathscr{Z}^{\vartheta,\phi}_{s}\left(p_{\varepsilon}(\cdot-z)\right)^{2}\psi(z)^{2}\text{\rm d}z\text{\rm d}s\quad\text{uniformly on $[0,T]$ in probability}

for any T0T\geq 0.

Also, we give a corollary on the quadratic variation.

Corollary 3.13.

For any ϑ\vartheta\in{\mathbb{R}}, ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}), ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}), and t>0t>0

E[θ,ϕ(ψ)t]=4π0<u<v<tdudv2dx2dyΦu2(x)Gϑ(vu,yx)ψ(y)2.\displaystyle\displaystyle E\left[\left\langle\mathscr{M}^{\theta,\phi}(\psi)\right\rangle_{t}\right]={4\pi}\int_{0<u<v<t}\text{\rm d}u\text{\rm d}v\int_{{\mathbb{R}}^{2}}\text{\rm d}x\int_{{\mathbb{R}}^{2}}\text{\rm d}y\Phi_{u}^{2}(x)G_{\vartheta}(v-u,y-x)\psi(y)^{2}.
Proof.

Since we proved that MN,ϕ(ψ)Nt\left\langle M^{N,\phi}(\psi)\right\rangle_{Nt} is uniform integrable, we have

E[θ,ϕ(ψ)t]=limNE[MN,ϕ(ψ)Nt].\displaystyle E\left[\left\langle\mathscr{M}^{\theta,\phi}(\psi)\right\rangle_{t}\right]=\lim_{N\to\infty}E\left[\left\langle M^{N,\phi}(\psi)\right\rangle_{Nt}\right].

Then, we can use the same argument in the proof of Theorem 1.2 in [12]. ∎

3.3.1 Proof of Lemma 3.11

Since Cb2(2)C_{b}^{2}({\mathbb{R}}^{2}) is a separating set for MF(2)M_{F}({\mathbb{R}}^{2}),

n=nN𝖹N,ϕ,\displaystyle\mathcal{F}_{n}=\mathcal{F}_{\frac{n}{N}}^{\mathsf{Z}^{N,\phi}},

and t𝖹N,ϕ=nN𝖹N,ϕ\mathcal{F}_{t}^{\mathsf{Z}^{N,\phi}}=\mathcal{F}_{\frac{n}{N}}^{\mathsf{Z}^{N,\phi}} for nNt<n+1N\frac{n}{N}\leq t<\frac{n+1}{N}.

Let 0s1sns<t0\leq s_{1}\leq\dots\leq s_{n}\leq s<t, n1n\geq 1 and FF be a bounded continuous function on MF(2)nM_{F}({\mathbb{R}}^{2})^{n}. Then, we have

E[(MNtN,ϕ(ψ)MNsN,ϕ(ψ))F(𝖹s1N,ϕ,,𝖹snN,ϕ)]=0\displaystyle E\left[\left(M^{N,\phi}_{\left\lfloor{Nt}\right\rfloor}(\psi)-M^{N,\phi}_{\left\lfloor{Ns}\right\rfloor}(\psi)\right)F\left(\mathsf{Z}^{N,\phi}_{s_{1}},\dots,\mathsf{Z}^{N,\phi}_{s_{n}}\right)\right]=0

The uniform integrability of MN,ϕ(ψ)2M^{N,\phi}_{\cdot}(\psi)^{2} and MN,ϕ(ψ)\left\langle M^{N,\phi}(\psi)\right\rangle_{\cdot} implies

E[(tϑ,ϕ(ψ)sϑ,ϕ(ψ))F(𝒵s1ϑ,ϕ,,𝒵snϑ,ϕ)]=0.\displaystyle E\left[\left(\mathscr{M}^{\vartheta,\phi}_{t}(\psi)-\mathscr{M}^{\vartheta,\phi}_{s}(\psi)\right)F\left(\mathscr{Z}^{\vartheta,\phi}_{s_{1}},\dots,\mathscr{Z}^{\vartheta,\phi}_{s_{n}}\right)\right]=0.

Thus, we completed the proof of Lemma 3.11.

 

3.3.2 Proof of Lemma 3.12

Since we know that for each t0t\geq 0 and ε>0\varepsilon>0

0t2𝒵sϑ,ϕ(pε(z))2ψ(z)2dz=limN0NtN2𝖹¯sN,ϕ(pε(z))2ψ(z)2dzdsa.s.,\displaystyle\int_{0}^{t}\int_{{\mathbb{R}}^{2}}\mathscr{Z}^{\vartheta,\phi}_{s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z=\lim_{N\to\infty}\int_{0}^{\frac{\left\lfloor{Nt}\right\rfloor}{N}}\int_{{\mathbb{R}}^{2}}\overline{\mathsf{Z}}^{N,\phi}_{s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z\text{\rm d}s\quad\text{a.s.,}

Fatou’s lemma implies that

E[(4πlogε0t2𝒵sϑ,ϕ(pε(z))2ψ(z)2dzdsϑ,ϕ(ψ)t)2]\displaystyle E\left[\left(\frac{{4\pi}}{-\log\varepsilon}\int_{0}^{t}\int_{{\mathbb{R}}^{2}}\mathscr{Z}^{\vartheta,\phi}_{s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z\text{\rm d}s-\left\langle\mathscr{M}^{\vartheta,\phi}_{\cdot}(\psi)\right\rangle_{t}\right)^{2}\right]
lim¯NE[(0NtN(4πlogε2𝖹¯sN,ϕ(pε(z))2ψ(z)2dzσN2Ny2Z¯N;Nsϕ(y)2ψN(y)2)ds)2].\displaystyle\leq\varliminf_{N\to\infty}E\left[\left(\int_{0}^{\frac{\lfloor Nt\rfloor}{N}}\left(\frac{{4\pi}}{-\log\varepsilon}\int_{{\mathbb{R}}^{2}}\overline{\mathsf{Z}}^{N,\phi}_{s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z-{\frac{\sigma_{N}^{2}}{N}\sum_{y\in{\mathbb{Z}}^{2}}{\overline{Z}}_{N;\left\lfloor{Ns}\right\rfloor}^{\phi}(y)^{2}\psi_{N}(y)^{2}}\right)\text{\rm d}s\right)^{2}\right].

We will prove the following lemma.

Lemma 3.14.

We have

limε0lim¯NE[(0NtN(4πlogε2𝖹¯sN,ϕ(pε(z))2ψ(z)2dzσN2Ny2Z¯N;Nsϕ(y)2ψN(y)2)ds)2]=0\displaystyle\lim_{\varepsilon\to 0}\varliminf_{N\to\infty}E\left[\left(\int_{0}^{\frac{\lfloor Nt\rfloor}{N}}\left(\frac{{4\pi}}{-\log\varepsilon}\int_{{\mathbb{R}}^{2}}\overline{\mathsf{Z}}^{N,\phi}_{s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z-{\frac{\sigma_{N}^{2}}{N}\sum_{y\in{\mathbb{Z}}^{2}}{\overline{Z}}_{N;\left\lfloor{Ns}\right\rfloor}^{\phi}(y)^{2}\psi_{N}(y)^{2}}\right)\text{\rm d}s\right)^{2}\right]=0 (3.14)

for t0t\geq 0, ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}), and ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}).

Proof of Lemma 3.12.

Let D[0,T]D\subset[0,T] be a dense subset. Then, we can see from Lemma 3.14 that for any sequence {εn}n1\{\varepsilon_{n}\}_{n\geq 1} with εn0\varepsilon_{n}\to 0, there exists a subsequence {εnk}\{\varepsilon_{n_{k}}\} such that

4πlogεnk0t2𝒵sϑ,ϕ(pεnk(z))2ψ(z)2dzdsϑ,ϕ(ψ)t\displaystyle\frac{{4\pi}}{-\log\varepsilon_{n_{k}}}\int_{0}^{t}\int_{{\mathbb{R}}^{2}}\mathscr{Z}^{\vartheta,\phi}_{s}(p_{\varepsilon_{n_{k}}}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z\text{\rm d}s\to\left\langle\mathscr{M}^{\vartheta,\phi}_{\cdot}(\psi)\right\rangle_{t}

for tDt\in D a.s. Since both processes are continuous and non-decreasing, this is uniform convergence on [0,T][0,T] a.s., and hence, Lemma 3.12 follows. ∎

The proof of Lemma is postponed to Section 6.

We will verify the convergence of expectations of quadratic variation as an exercise and give the proof of existence of extension in Theorem 1.18: Theorem 2.3 and Lemma 3.6 imply that

E[ϑ,ϕ(ψ)t]\displaystyle E\left[\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}\right]
=limNE[MN,ϕ(ψ)Nt]=4π0<u<v<tdudv2×2dxdyΦu(x)2Gϑ(vu,yx)ψ(y)2.\displaystyle=\lim_{N\to\infty}E\left[\left\langle M^{N,\phi}(\psi)\right\rangle_{Nt}\right]=4\pi\int_{0<u<v<t}\text{\rm d}u\text{\rm d}v\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}x\text{\rm d}y\Phi_{u}(x)^{2}G_{\vartheta}(v-u,y-x)\psi(y)^{2}. (3.15)

Also,

limNE[0NtN4πlogε2𝖹¯sN,ϕ(pε(z))2ψ(z)2dz]\displaystyle\lim_{N\to\infty}E\left[\int_{0}^{\frac{\lfloor Nt\rfloor}{N}}\frac{4\pi}{-\log\varepsilon}\int_{{\mathbb{R}}^{2}}\overline{\mathsf{Z}}^{N,\phi}_{s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z\right]
=0t4πlogε2(2dxϕ(x)pε(xy)dx)2ψ(y)2dyds\displaystyle=\int_{0}^{t}\frac{4\pi}{-\log\varepsilon}\int_{{\mathbb{R}}^{2}}\left(\int_{{\mathbb{R}}^{2}}\text{\rm d}x\phi(x)p_{\varepsilon}(x-y)\text{\rm d}x\right)^{2}\psi(y)^{2}\text{\rm d}y\text{\rm d}s
+0tds16π2logε0<u<v<sdudv2×2dydzΦu(x)2Gϑ(vu,yx)psv+ε(zy)2ψ(z)2,\displaystyle+\int_{0}^{t}\text{\rm d}s\frac{16\pi^{2}}{-\log\varepsilon}\int_{0<u<v<s}\text{\rm d}u\text{\rm d}v\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}y\text{\rm d}z\Phi_{u}(x)^{2}G_{\vartheta}(v-u,y-x)p_{{s-v+\varepsilon}}(z-y)^{2}\psi(z)^{2},

where we have used that 2pvu(yw)pε(wz)dw=pvu+ε(yz)\int_{{\mathbb{R}}^{2}}p_{v-u}(y-w)p_{\varepsilon}(w-z)\text{\rm d}w=p_{v-u+\varepsilon}(y-z). Since pt(x)2=14πtpt2(x)p_{t}(x)^{2}=\frac{1}{4\pi t}p_{\frac{t}{2}}(x), we have

limε0limNE[0NtN4πlogε2𝖹¯N;sϕ(pε(z))2ψ(z)2dz]\displaystyle\lim_{\varepsilon\to 0}\lim_{N\to\infty}E\left[\int_{0}^{\frac{\lfloor Nt\rfloor}{N}}\frac{4\pi}{-\log\varepsilon}\int_{{\mathbb{R}}^{2}}\overline{\mathsf{Z}}^{\phi}_{N;s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z\right]
=limε00<u<v<tdudvvtds16π2logε2dz2dyΦu(x)2Gϑ(vu,yx)14π(sv+ε)psv+ε2(zx)ψ(z)2\displaystyle=\lim_{\varepsilon\to 0}\int_{0<u<v<t}\text{\rm d}u\text{\rm d}v\int_{v}^{t}\text{\rm d}s\frac{16\pi^{2}}{-\log\varepsilon}\int_{{\mathbb{R}}^{2}}\text{\rm d}z\int_{{\mathbb{R}}^{2}}\text{\rm d}y\Phi_{u}(x)^{2}G_{\vartheta}(v-u,y-x)\frac{1}{4\pi(s-v+\varepsilon)}p_{\frac{s-v+\varepsilon}{2}}(z-x)\psi(z)^{2}
=4π0<u<v<tdudv2dz2dyΦu(x)2Gϑ(vu,yx)ψ(z)2,\displaystyle=4\pi\int_{0<u<v<t}\text{\rm d}u\text{\rm d}v\int_{{\mathbb{R}}^{2}}\text{\rm d}z\int_{{\mathbb{R}}^{2}}\text{\rm d}y\Phi_{u}(x)^{2}G_{\vartheta}(v-u,y-x)\psi(z)^{2},

where we have used the following lemma and the dominated convergence theorem in the last equation.

Lemma 3.15.

Let ψb(2)\psi\in\mathcal{B}_{b}({\mathbb{R}}^{2}) and T>0T>0. Then, for each 0<t<T0<t<T, there exists CT,ψC_{T,\psi} such that

sup0<ε<12|1logε2dz0tds1s+εps+2ε2(zx)ψ(z)|CT,ψ\displaystyle\sup_{0<\varepsilon<\frac{1}{2}}\left|\frac{-1}{\log\varepsilon}\int_{{\mathbb{R}}^{2}}\text{\rm d}z\int_{0}^{t}\text{\rm d}s\frac{1}{s+\varepsilon}p_{\frac{s+2\varepsilon}{2}}(z-x)\psi(z)\right|\leq C_{T,\psi} (3.16)
for each x2x\in{\mathbb{R}}^{2} and
limε01logε2dz0tds1s+εps+ε2(zx)ψ(z)=ψ(x)\displaystyle\lim_{\varepsilon\to 0}\frac{-1}{\log\varepsilon}\int_{{\mathbb{R}}^{2}}\text{\rm d}z\int_{0}^{t}\text{\rm d}s\frac{1}{s+\varepsilon}p_{\frac{s+\varepsilon}{2}}(z-x)\psi(z)=\psi(x) (3.17)

for a.e. xx for 0<tT0<t\leq T.

Proof.

It is easy to see that

|1logε2dz0tds1s+εps+ε2(zx)ψ(z)|\displaystyle\left|\frac{-1}{\log\varepsilon}\int_{{\mathbb{R}}^{2}}\text{\rm d}z\int_{0}^{t}\text{\rm d}s\frac{1}{s+\varepsilon}p_{\frac{s+\varepsilon}{2}}(z-x)\psi(z)\right|
ψ1logε[log(t+ε)log(ε)]CTψ\displaystyle\leq\|\psi\|_{\infty}\frac{-1}{\log\varepsilon}\left[\log(t+\varepsilon)-\log(\varepsilon)\right]\leq C_{T}\|\psi\|_{\infty}

for some CT>0C_{T}>0.

Also,

2dz0tds1s+εps+ε2(zx)ψ(z)\displaystyle\int_{{\mathbb{R}}^{2}}\text{\rm d}z\int_{0}^{t}\text{\rm d}s\frac{1}{s+\varepsilon}p_{\frac{s+\varepsilon}{2}}(z-x)\psi(z)
=2dzψ(z)1π|zx|2(exp(|zx|2t+ε)exp(|zx|2ε))\displaystyle=\int_{{\mathbb{R}}^{2}}\text{\rm d}z\psi(z)\frac{1}{\pi|z-x|^{2}}\left(\exp\left(-\frac{|z-x|^{2}}{t+\varepsilon}\right)-\exp\left(-\frac{|z-x|^{2}}{\varepsilon}\right)\right)
=0dr[0,2π]dθψ(x+r(cosθ,sinθ))1πr(exp(r2t+ε)exp(r2ε)).\displaystyle=\int_{0}^{\infty}\text{\rm d}r\int_{[0,2\pi]}\text{\rm d}\theta\psi\left(x+r(\cos\theta,\sin\theta)\right)\frac{1}{\pi r}\left(\exp\left(-\frac{r^{2}}{t+\varepsilon}\right)-\exp\left(-\frac{r^{2}}{\varepsilon}\right)\right).

Therefore, we obtain from l’Hôpital’s rule that

limε01logε2dz0tds1s+εps+ε2(zx)ψ(z)\displaystyle\lim_{\varepsilon\to 0}\frac{1}{-\log\varepsilon}\int_{{\mathbb{R}}^{2}}\text{\rm d}z\int_{0}^{t}\text{\rm d}s\frac{1}{s+\varepsilon}p_{\frac{s+\varepsilon}{2}}(z-x)\psi(z)
=limε0ε0dr[0,2π]dθψ(x+r(cosθ,sinθ))1πrddε(exp(r2t+ε)exp(r2ε))\displaystyle=-\lim_{\varepsilon\to 0}\varepsilon\int_{0}^{\infty}\text{\rm d}r\int_{[0,2\pi]}\text{\rm d}\theta\psi\left(x+r(\cos\theta,\sin\theta)\right)\frac{1}{\pi r}\frac{\text{\rm d}}{\text{\rm d}\varepsilon}\left(\exp\left(-\frac{r^{2}}{t+\varepsilon}\right)-\exp\left(-\frac{r^{2}}{\varepsilon}\right)\right)

if the limit in the righthand side exists for each t>0t>0.

It is easy to see that it should be equal to

limε00dr[0,2π]dθψ(x+r(cosθ,sinθ))rπεexp(r2ε)=limε02pε(zx)ψ(z)dz=ψ(x)\displaystyle\lim_{\varepsilon\to 0}\int_{0}^{\infty}\text{\rm d}r\int_{[0,2\pi]}\text{\rm d}\theta\psi\left(x+r(\cos\theta,\sin\theta)\right)\frac{r}{\pi\varepsilon}\exp\left(-\frac{r^{2}}{\varepsilon}\right)=\lim_{\varepsilon\to 0}\int_{{\mathbb{R}}^{2}}p_{\varepsilon}(z-x)\psi(z)\text{\rm d}z=\psi(x)

a.e. xx by Lebesgue’s differential theorem. Therefore, (3.17) holds for ψb(2)\psi\in\mathcal{B}_{b}({\mathbb{R}}^{2}). ∎

Proof of existence of extension in Theorem 1.18.

Let ψ\psi be a bounded Borel function. Then, Lusin’s theorem and Tietze extension theorem implies that there exists a sequence {ψ^n}\{\hat{\psi}_{n}\} in Cb(2)C_{b}({\mathbb{R}}^{2}) such that

ψ^n(x)ψ(x)a.e. x and ψ^n=ψ for all n1.\displaystyle\hat{\psi}_{n}(x)\to\psi(x)\quad\text{a.e.~{}$x$ and }\|\hat{\psi}_{n}\|_{\infty}=\|\psi\|_{\infty}\text{ for all $n\geq 1$.}

(See [43, Remark 1.3.30].) Also, we know any bounded continuous function ff can be approximated by pϵfp_{\epsilon}*f uniformly on any compact sets.

Hence, we can choose ϵn\epsilon_{n} such that

pϵnψ^n(x)ψ(x)a.e. x and pϵnψ^nψ for all n1.\displaystyle p_{\epsilon_{n}}*\hat{\psi}_{n}(x)\to\psi(x)\quad\text{a.e.~{}$x$ and }\|p_{\epsilon_{n}}*\hat{\psi}_{n}\|_{\infty}\leq\|\psi\|_{\infty}\text{ for all $n\geq 1$.} (3.18)

For each ψb(2)\psi\in\mathcal{B}_{b}({\mathbb{R}}^{2}), we set ψn(x)=pϵnψ^n(x)\psi_{n}(x)=p_{\epsilon_{n}}*\hat{\psi}_{n}(x) satifying (3.18). Then, we can see from the Burkholder-Davis-Gundy inequality and (3.15) that for n,m1n,m\geq 1, and t0t\geq 0

E[sup0st(sϑ,ϕ(ψn)sϑ,ϕ(ψm))2]\displaystyle E\left[\sup_{0\leq s\leq t}\left(\mathscr{M}_{s}^{\vartheta,\phi}(\psi_{n})-\mathscr{M}_{s}^{\vartheta,\phi}(\psi_{m})\right)^{2}\right] CE[ϑ,ϕ(ψnψm)t]\displaystyle\leq CE\left[\left\langle\mathscr{M}^{\vartheta,\phi}(\psi_{n}-\psi_{m})\right\rangle_{t}\right]
=4πC0<u<v<tdudv2×2dxdyΦu(x)2Gϑ(vu,yx)(ψn(y)ψm(y))2\displaystyle=4\pi C\int_{0<u<v<t}\text{\rm d}u\text{\rm d}v\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}x\text{\rm d}y\Phi_{u}(x)^{2}G_{\vartheta}(v-u,y-x)\left(\psi_{n}(y)-\psi_{m}(y)\right)^{2} (3.19)

for some constant C>0C>0. Thus, the dominated convergence theorem implies that {ϑ,ϕ(ψn)}n1\displaystyle\left\{\mathscr{M}_{\cdot}^{\vartheta,\phi}(\psi_{n})\right\}_{n\geq 1} is an L2L^{2}-Cauchy sequence and CC-tight and hence the limit exists and we denote it by ϑ,ϕ(ψ)\mathscr{M}^{\vartheta,\phi}_{\cdot}(\psi).

The proof for the representation of the quadratic variations of tϑ,ϕ(ψ)\mathscr{M}^{\vartheta,\phi}_{t}(\psi) will be given in Section 6. However, the above proof implies the following convergence of the quadratic variation.

Corollary 3.16.

Let ϑ\vartheta\in{\mathbb{R}}, ϕCc2(2)\phi\in C_{c}^{2}({\mathbb{R}}^{2}). Then, for each ψb(2)\psi\in\mathcal{B}_{b}({\mathbb{R}}^{2}) and t>0t>0,

limnE[sup0st|ϑ,ϕ(ψ)sϑ,ϕ(ψn)s|]=0,\displaystyle\lim_{n\to\infty}E\left[\sup_{0\leq s\leq t}\left|\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{s}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi_{n})\right\rangle_{s}\right|\right]=0,

where {ψn}\{\psi_{n}\} is a sequence in Cb2(2)C_{b}^{2}({\mathbb{R}}^{2}) such that ψn(x)\psi_{n}(x) converges to ψ(x)\psi(x) for any x2x\in{\mathbb{R}}^{2} and ψnψ\|\psi_{n}\|_{\infty}\leq\|\psi\|_{\infty}.

Proof.

From (1.11), we can see that for n,m1n,m\geq 1

|ϑ,ϕ(ψn)sϑ,ϕ(ψm)s|ϑ,ϕ(ψn+ψm)t12ϑ,ϕ(ψnψm)t12.\displaystyle\left|\left\langle\mathscr{M}^{\vartheta,\phi}(\psi_{n})\right\rangle_{s}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi_{m})\right\rangle_{s}\right|\leq\left\langle\mathscr{M}^{\vartheta,\phi}(\psi_{n}+\psi_{m})\right\rangle_{t}^{\frac{1}{2}}\left\langle\mathscr{M}^{\vartheta,\phi}(\psi_{n}-\psi_{m})\right\rangle_{t}^{\frac{1}{2}}.

for 0st0\leq s\leq t a.s. Then,

E[sup0st|ϑ,ϕ(ψn)sϑ,ϕ(ψm)s|]E[ϑ,ϕ(ψn+ψm)t]12E[ϑ,ϕ(ψnψm)t]12.\displaystyle E\left[\sup_{0\leq s\leq t}\left|\left\langle\mathscr{M}^{\vartheta,\phi}(\psi_{n})\right\rangle_{s}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi_{m})\right\rangle_{s}\right|\right]\leq E\left[\left\langle\mathscr{M}^{\vartheta,\phi}(\psi_{n}+\psi_{m})\right\rangle_{t}\right]^{\frac{1}{2}}E\left[\left\langle\mathscr{M}^{\vartheta,\phi}(\psi_{n}-\psi_{m})\right\rangle_{t}\right]^{\frac{1}{2}}. (3.20)

The same argument in the proof of Theorem 1.18 implies the righthand side converges to zero. Thus, {ϑ,ϕ(ϕn)}n1\displaystyle\left\{\left\langle\mathscr{M}^{\vartheta,\phi}(\phi_{n})\right\rangle_{\cdot}\right\}_{n\geq 1} is a L2L^{2}-Cauchy sequence and CC-tight. So, the limit denoted by ϑ,ϕ(ϕ)\left\langle\mathscr{M}^{\vartheta,\phi}(\phi)\right\rangle_{\cdot} is the quadratic variation of ϑ,ϕ(ϕ)\mathscr{M}^{\vartheta,\phi}_{\cdot}(\phi). The statement follows by taking limit mm\to\infty in (3.20). ∎

Remark 3.17.

Combining Theorem 2.3 and Doob’s inequality, we can find that the sequence of process {𝒵ϑ,ϕ(ψn)}\{\mathscr{Z}^{\vartheta,\phi}_{\cdot}(\psi_{n})\} weakly converges to a process 𝒵ϑ,ϕ(ψ)\mathscr{Z}^{\vartheta,\phi}_{\cdot}(\psi).

The Skorkhod representation theorem allows us to 𝒵tϑ,ϕ(ψ)\mathscr{Z}^{\vartheta,\phi}_{t}(\psi) has the same form (1.10).

Remark 3.18.

We remark that every term in expectation in (3.9) and (3.14) can be described via partition functions from (3.4). More precisely, we need look at

σN4N2[s,t]2dudvy1,y2E[Z¯N;uϕ(y1)2Z¯N;vϕ(y2)2]\displaystyle\frac{\sigma_{N}^{4}}{N^{2}}\int_{[s,t]^{2}}\text{\rm d}u\text{\rm d}v\sum_{y^{1},y^{2}}E\left[\overline{Z}_{N;u}^{\phi}(y^{1})^{2}\overline{Z}_{N;v}^{\phi}(y^{2})^{2}\right] (3.21)

for Lemma 3.7, and the linear combination of

σN4N2[0,t]2dudvy1,y2E[Z¯N;uϕ(y1)2Z¯N;vϕ(y2)2]ψN(y1)2ψN(y2)2\displaystyle\frac{\sigma_{N}^{4}}{N^{2}}\int_{[0,t]^{2}}\text{\rm d}u\text{\rm d}v\sum_{y^{1},y^{2}}E\left[\overline{Z}_{N;u}^{\phi}(y^{1})^{2}\overline{Z}_{N;v}^{\phi}(y^{2})^{2}\right]\psi_{N}(y^{1})^{2}\psi_{N}(y^{2})^{2} (3.22)
σN2N3[0,t]2dudvy1,y2,y3E[Z¯N;uϕ(y1)Z¯N;uϕ(y2)Z¯N;vϕ(y3)2]\displaystyle\frac{\sigma_{N}^{2}}{N^{3}}\int_{[0,t]^{2}}\text{\rm d}u\text{\rm d}v\sum_{y^{1},y^{2},y^{3}}E\left[\overline{Z}_{N;u}^{\phi}(y^{1})\overline{Z}_{N;u}^{\phi}(y^{2})\overline{Z}_{N;v}^{\phi}(y^{3})^{2}\right]
×2dzpε(zy1N)pε(zy2N)ψ(z)2ψN(y3)2\displaystyle\hskip 30.00005pt\times\int_{{\mathbb{R}}^{2}}\text{\rm d}zp_{\varepsilon}\left(z-\frac{y^{1}}{\sqrt{N}}\right)p_{\varepsilon}\left(z-\frac{y^{2}}{\sqrt{N}}\right)\psi(z)^{2}\psi_{N}(y^{3})^{2} (3.23)
1N4[0,t]2dudvy1,y2,y3,y4E[Z¯N;uϕ(y1)Z¯N;uϕ(y2)Z¯N;vϕ(y3)Z¯N;vϕ(y4)]\displaystyle\frac{1}{N^{4}}\int_{[0,t]^{2}}\text{\rm d}u\text{\rm d}v\sum_{y^{1},y^{2},y^{3},y^{4}}E\left[\overline{Z}_{N;u}^{\phi}(y^{1})\overline{Z}_{N;u}^{\phi}(y^{2})\overline{Z}_{N;v}^{\phi}(y^{3})\overline{Z}_{N;v}^{\phi}(y^{4})\right]
×(2)2dz1dz2pε(z1y1N)pε(z1y2N)pε(z2y3N)pε(z2y4N)ψ(z1)2ψN(z2)2.\displaystyle\hskip 30.00005pt\times\int_{({\mathbb{R}}^{2})^{2}}\text{\rm d}z_{1}\text{\rm d}z_{2}p_{\varepsilon}\left(z_{1}-\frac{y^{1}}{\sqrt{N}}\right)p_{\varepsilon}\left(z_{1}-\frac{y^{2}}{\sqrt{N}}\right)p_{\varepsilon}\left(z_{2}-\frac{y^{3}}{\sqrt{N}}\right)p_{\varepsilon}\left(z_{2}-\frac{y^{4}}{\sqrt{N}}\right)\psi(z_{1})^{2}\psi_{N}(z_{2})^{2}. (3.24)

for Lemma 3.14.

Thus, we will entirely focused on computing of moments of partition functions in the following sections.

4 Moments of partition functions

From now, we will omit the parameter NN in the notations if it is clear from the context.

4.1 Chaos expansion and moments

Let 𝕋\mathbb{T} be a countable set and {ωt}t𝕋\{\omega_{t}\}_{t\in\mathbb{T}} be independent Bernoulli distributed random variables with P(ωt=1)=P(ωt=1)=12P(\omega_{t}=1)=P(\omega_{t}=-1)=\frac{1}{2}.

For a finite subset F𝕋F\subset\mathbb{T}, we define

ωF=tFωt\displaystyle\omega_{F}=\prod_{t\in F}\omega_{t}

and the polynomial chaos P(ω)P(\omega) is defined as a linear combination of {ωF:F𝕋 is finite}\{\omega_{F}:F\subset\mathbb{T}\text{ is finite}\}, i.e.

P(ω)=F𝕋:finiteaFωF\displaystyle P(\omega)=\sum_{F\subset\mathbb{T}:\text{finite}}a_{F}\omega_{F}

for some {aF}F𝕋:finite\{a_{F}\}_{F\subset\mathbb{T}:\text{finite}} such that aF=0a_{F}=0 except for some F1,,FmP𝕋F_{1},\dots,F_{m_{P}}\subset\mathbb{T}. We set ω=1\omega_{\emptyset}=1 for convention.

Then, it is easy to see that 𝔼[ωF1ωFk]=1\mathbb{E}\left[\omega_{F_{1}}\dots\omega_{F_{k}}\right]=1 if any tF1Fkt\in F_{1}\cup\dots\cup F_{k} belongs to exactly even number of subsets Fk1,,Fk2lpF_{k_{1}},\dots,F_{k_{2l_{p}}}, and is equal to 0 otherwise.

Now, we will give the polynomial chaos expansion of partition functions: Take 𝕋=3\mathbb{T}={\mathbb{Z}}^{3}. For t>0t>0, y3y\in{\mathbb{Z}}^{3}, ϕCc(2)\phi\in C_{c}({\mathbb{R}}^{2}),

Z¯N;tϕ(y)\displaystyle\overline{Z}_{N;t}^{\phi}(y)
=qNt(ϕN,y)+k=11n1<<nk<Ntx1,,xk2ξn1,x1βqn1(ϕN,x1)(i=2kξni,xiβqni1,ni(xi1,xi))qnk,Nt(xk,y)\displaystyle=q_{Nt}(\phi_{N},y)+\sum_{k=1}^{\infty}\sum_{\begin{smallmatrix}1\leq n_{1}<\dots<n_{k}<Nt\\ x_{1},\dots,x_{k}\in{\mathbb{Z}}^{2}\end{smallmatrix}}\xi_{n_{1},x_{1}}^{\beta}q_{n_{1}}(\phi_{N},x_{1})\left(\prod_{i=2}^{k}\xi_{n_{i},x_{i}}^{\beta}q_{n_{i-1},n_{i}}(x_{i-1},x_{i})\right)q_{n_{k},Nt}(x_{k},y)
=|A|=0𝐀3a𝐀(t)(ϕ,y)ξ𝐀β,\displaystyle=\sum_{|A|=0}^{\infty}\sum_{\mathbf{A}\subset{\mathbb{Z}}^{3}}a_{\mathbf{A}}^{(t)}(\phi,y)\xi_{\mathbf{A}}^{\beta},

where for 𝐀=(A1,,A|𝐀|){\mathbf{A}}=(A_{1},\dots,A_{|\mathbf{A}|}) with Ai=(ni,xi)3A_{i}=(n_{i},x_{i})\in{\mathbb{Z}}^{3} (1n1<<n|𝐀|<Nt1\leq n_{1}<\dots<n_{|\mathbf{A}|}<Nt), we define

qn(ϕN,y)=x2ϕN(x)qn(x,y)for n0y2\displaystyle q_{n}(\phi_{N},y)=\sum_{x\in{\mathbb{Z}}^{2}}\phi_{N}(x)q_{n}(x,y)\quad\text{for $n\geq 0$, $y\in{\mathbb{Z}}^{2}$}
ξ𝐀β=(n,x)𝐀ξn,xβ\displaystyle\xi_{\mathbf{A}}^{\beta}=\prod_{(n,x)\in{\mathbf{A}}}\xi_{n,x}^{\beta}

and

a𝐀(t)(ϕ,y)={qNt(ϕN,y)if 𝐀=qn1(ϕN,x1)(i=2|𝐀|q(Ai1,Ai))q(n|𝐀|,n|𝐀|+1)if 𝐀=(A1,,A|𝐀|)0otherwise.\displaystyle a_{\mathbf{A}}^{(t)}(\phi,y)=\begin{cases}q_{Nt}(\phi_{N},y)\quad&\text{if }{\mathbf{A}}=\emptyset\\ q_{n_{1}}(\phi_{N},x_{1})\left(\prod_{i=2}^{|{\mathbf{A}}|}q(A_{i-1},A_{i})\right)q(n_{|{\mathbf{A}}|},n_{|{\mathbf{A}}|+1})\quad&\text{if }\begin{array}[]{l}{\mathbf{A}}=(A_{1},\dots,A_{|{\mathbf{A}}|})\end{array}\\ 0\quad&\text{otherwise}.\end{cases}

Here, we define i=21()=1\prod_{i=2}^{1}(\cdots)=1 for our convention and we set

A|𝐀|+1=(Nt,y)\displaystyle A_{|\mathbf{A}|+1}=(Nt,y)
q(Ai1,Ai)=qnini1(xi1,xi).\displaystyle q(A_{i-1},A_{i})=q_{n_{i}-n_{i-1}}(x_{i-1},x_{i}).

For our convenience, we introduce a set of finite subsets of indices

𝙵(T)=𝙵T\displaystyle\mathtt{F}(T)=\mathtt{F}_{T}
:={𝐀3:𝐀=(A1,,A|𝐀|),Ai=(ni,xi)3,1n1<<n|𝐀|<NT}\displaystyle:=\left\{\mathbf{A}\subset{\mathbb{Z}}^{3}:\begin{array}[]{l}\mathbf{A}=(A_{1},\dots,A_{|\mathbf{A}|}),\quad A_{i}=(n_{i},x_{i})\in{\mathbb{Z}}^{3},\\ 1\leq n_{1}<\dots<n_{|\mathbf{A}|}<NT\end{array}\right\}

for T0T\geq 0, where we set |𝐀|=0|\mathbf{A}|=0 for 𝐀=\mathbf{A}=\emptyset. We define 𝙵=T0𝙵T\displaystyle\mathtt{F}=\bigcup_{T\geq 0}\mathtt{F}_{T}.

The above argument gives that

𝔼[Z¯N;sϕ(ya)Z¯N;sϕ(yb)Z¯N;tϕ(yc)Z¯N;tϕ(yd)]\displaystyle\mathbb{E}\left[\overline{Z}_{N;s}^{\phi}(y^{a})\overline{Z}_{N;s}^{\phi}(y^{b})\overline{Z}_{N;t}^{\phi}(y^{c})\overline{Z}_{N;t}^{\phi}(y^{d})\right]
=𝐀,𝐁𝙵(s)𝐂,𝐃𝙵(t)a𝐀(s)(ϕ,ya)a𝐁(s)(ϕ,yb)a𝐂(t)(ϕ,yc)a𝐃(t)(ϕ,yd)𝔼[ξ𝐀βNξ𝐁βNξ𝐂βNξ𝐃βN].\displaystyle=\sum_{\begin{smallmatrix}\mathbf{A},\mathbf{B}\in\mathtt{F}({s})\end{smallmatrix}}\sum_{\begin{smallmatrix}\mathbf{C},\mathbf{D}\in\mathtt{F}(t)\end{smallmatrix}}a_{\mathbf{A}}^{(s)}(\phi,y^{a})a_{\mathbf{B}}^{(s)}(\phi,y^{b})a_{\mathbf{C}}^{(t)}(\phi,y^{c})a_{\mathbf{D}}^{(t)}(\phi,y^{d})\mathbb{E}\left[\xi_{\mathbf{A}}^{\beta_{N}}\xi_{\mathbf{B}}^{\beta_{N}}\xi_{\mathbf{C}}^{\beta_{N}}\xi_{\mathbf{D}}^{\beta_{N}}\right]. (4.1)

Now, we focus on the finite subsets 𝐀,𝐁,𝐂,𝐃𝙵\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D}\in\mathtt{F} that contribute to the summation in the right-hand side.

We say 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} have an odd intersection if one of the following holds:

  1. (1)

    there exists an (n,x)(n,x) such that (n,x)(n,x) belongs to three of 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} but does not belong to the other one.

  2. (2)

    there exists an (n,x)(n,x) such that (n,x)(n,x) belongs to one of 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} but does not belong to the others.

Also, we say 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} have even intersections if 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} don’t have an odd intersection.

We have from (3.2)

𝔼[ξ𝐀βNξ𝐁βNξ𝐂βNξ𝐃βN]={σN|𝐀|+|𝐁|+|𝐂|+|𝐃|if 𝐀,𝐁,𝐂,𝐃:even intersection0if 𝐀,𝐁,𝐂,𝐃:even intersection\displaystyle\mathbb{E}\left[\xi_{\mathbf{A}}^{\beta_{N}}\xi_{\mathbf{B}}^{\beta_{N}}\xi_{\mathbf{C}}^{\beta_{N}}\xi_{\mathbf{D}}^{\beta_{N}}\right]=\begin{cases}\sigma_{N}^{|\mathbf{A}|+|\mathbf{B}|+|\mathbf{C}|+|\mathbf{D}|}\quad&\text{if }\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D}:\text{even intersection}\\ 0\quad&\text{if }\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D}:\text{even intersection}\\ \end{cases}

Thus, 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} that have even intersections contribute to the summation of 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} in (4.1).


The above argument yields that

(3.21)\displaystyle\eqref{eq:4thmomentQV} =σN4N4NsNuNtNsNvNty1,y2𝐀,𝐁𝙵(u)𝐂,𝐃𝙵(v)even intersectionsσN|𝐀|++|𝐃|a𝐀(u)(ϕ,y1)a𝐁(u)(ϕ,y1)a𝐂(v)(ϕ,y2)a𝐃(v)(ϕ,y2)\displaystyle=\frac{\sigma_{N}^{4}}{N^{4}}\sum_{\begin{smallmatrix}Ns\leq Nu\leq Nt\\ Ns\leq Nv\leq Nt\end{smallmatrix}}\sum_{y^{1},y^{2}}\sum_{\begin{smallmatrix}\mathbf{A},\mathbf{B}\in\mathtt{F}(u)\\ \mathbf{C},\mathbf{D}\in\mathtt{F}(v)\\ \text{even intersections}\end{smallmatrix}}\sigma_{N}^{|\mathbf{A}|+\dots+|\mathbf{D}|}a^{(u)}_{\mathbf{A}}(\phi,y^{1})a^{(u)}_{\mathbf{B}}(\phi,y^{1})a^{(v)}_{\mathbf{C}}(\phi,y^{2})a^{(v)}_{\mathbf{D}}(\phi,y^{2}) (4.2)

We can write (3.22)-(3.24) in similar ways, but we omit giving them here.

Hereafter, we may assume that 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} have even intersection.

4.2 Pairings of intersections

We write elements of 𝐀𝐁𝐂𝐃\mathbf{A}\cup\mathbf{B}\cup\mathbf{C}\cup\mathbf{D} in time ordered as

𝐀𝐁𝐂𝐃={{(ni,xi)}i=1,,k:1n1nk},\displaystyle\mathbf{A}\cup\mathbf{B}\cup\mathbf{C}\cup\mathbf{D}=\{\{(n_{i},x_{i})\}_{i=1,\dots,k}:1\leq n_{1}\leq\dots\leq n_{k}\}, (4.3)

where for the case ni=ni+1n_{i}=n_{i+1}, we may choose xixi+1x_{i}\not=x_{i+1} such that (ni,xi)𝐀(n_{i},x_{i})\in\mathbf{A} since if ni=ni+1n_{i}=n_{i+1} for some ii, then (ni,xi)(n_{i},x_{i}) belong to two of 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} and (ni+1,xi+1)(n_{i+1},x_{i+1}) belongs to the other two). Also, we define by

𝐀𝐁𝐂𝐃|={1m1<m2<<ml:{n1,,nk}={m1,,ml}}\displaystyle\mathbf{A}\cup\mathbf{B}\cup\mathbf{C}\cup\mathbf{D}|_{\mathbb{N}}=\{1\leq m_{1}<m_{2}<\dots<m_{l}:\{n_{1},\dots,n_{k}\}=\{m_{1},\dots,m_{l}\}\}

the sequence of intersection times.

Definition 4.1.

Suppose 𝐀,𝐁,𝐂,𝐃𝙵\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D}\in\mathtt{F} with (4.3). If 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} have even intersections and 𝐀𝐁𝐂𝐃\mathbf{A}\cup\mathbf{B}\cup\mathbf{C}\cup\mathbf{D}\not=\emptyset, for each (ni,xi)𝐀𝐁𝐂𝐃(n_{i},x_{i})\in\mathbf{A}\cup\mathbf{B}\cup\mathbf{C}\cup\mathbf{D}, one of the following three cases occurs:

  1. (P1)(\mathrm{P}1)

    (ninjn_{i}\not=n_{j} for jij\not=i)

    1. (P1-i)(\mathrm{P}\text{$1$-}\mathrm{i})

      (ni,xi)(n_{i},x_{i}) belongs to two of 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} but not to the other two. Moreover, (ni,y)𝐀𝐁𝐂𝐃(n_{i},y)\not\in\mathbf{A}\cup\mathbf{B}\cup\mathbf{C}\cup\mathbf{D} for any yxy\not=x.

    2. (P1-ii)(\mathrm{P}\text{$1$-}\mathrm{ii})

      (ni,xi)𝐀𝐁𝐂𝐃(n_{i},x_{i})\in\mathbf{A}\cap\mathbf{B}\cap\mathbf{C}\cap\mathbf{D}.

  2. (P2)(\mathrm{P}2)

    (ni=nj=nn_{i}=n_{j}=n for iji\not=j) |ji|=1|j-i|=1 and (n,x)(n,x) belongs to two of 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} and (n,y)(n,y) belongs to the other two.

Thus, when 𝐀𝐁𝐂𝐃\mathbf{A}\cup\mathbf{B}\cup\mathbf{C}\cup\mathbf{D} has even intersections, each ni𝐀𝐁𝐂𝐃|n_{i}\in\mathbf{A}\cup\mathbf{B}\cup\mathbf{C}\cup\mathbf{D}|_{\mathbb{N}} has an associated pair(s) of indices, denoted by 𝓅i\mathscr{p}_{i}, EF{AB,AC,AD,BC,BD,CD}EF\in\{AB,AC,AD,BC,BD,CD\} (\leftrightarrow(P1)(\mathrm{P}1)), ABCDABCD (\leftrightarrow(P1)(\mathrm{P}1)(P1-𝑖𝑖)(\mathrm{P}\text{$1$-}\mathrm{ii})), [EF][GH][EF][GH] (\leftrightarrow(P2)(\mathrm{P}2)), where {E,F,G,H}={A,B,C,D}\{E,F,G,H\}=\{A,B,C,D\}.

We denote by 𝒫1\mathcal{P}_{1}, 𝒫2\mathcal{P}_{2} the set of pairs with type (P1)(\mathrm{P}1), with type (P2)(\mathrm{P}2), respectively and 𝒫:={ABCD}\mathcal{P}_{*}:=\{ABCD\}.

We set 𝒫=𝒫1𝒫2𝒫\mathcal{P}=\mathcal{P}_{1}\cup\mathcal{P}_{2}\cup\mathcal{P}_{*}. Then, we define the associated map ι\iota which maps 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} to a finite 𝒫\mathcal{P}-sequence if 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} have even intersection and 𝐀𝐁𝐂𝐃\mathbf{A}\cup\mathbf{B}\cup\mathbf{C}\cup\mathbf{D}\not=\emptyset:

ι(𝐀,𝐁,𝐂,𝐃)=(𝓅1,,𝓅l)=:𝐩.\displaystyle\iota(\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D})=(\mathscr{p}_{1},\dots,\mathscr{p}_{l})=:\mathbf{p}. (4.4)
Definition 4.2.

We say p=EF,q=GH𝒫1p=EF,q=GH\in\mathcal{P}_{1} are a couple if {E,F,G,H}={A,B,C,D}\{E,F,G,H\}=\{A,B,C,D\}, i.e. each of (AB,CD)(AB,CD), (AC,BD)(AC,BD), and (AD,BC)(AD,BC) is a couple.

We denote by 𝐏f\mathbf{P}_{f} the set of finite sequences of 𝒫\mathcal{P}. Then, (4.2) is rewritten by

σN4N4NsNuNtNsNvNty1,y2𝐩𝐏f𝐀,𝐁𝙵(u)𝐂,𝐃𝙵(v)even intersectionsι(𝐀,𝐁,𝐂,𝐃)=𝐩σN|𝐀|++|𝐃|a𝐀(u)(ϕ,y1)a𝐁(u)(ϕ,y1)a𝐂(v)(ϕ,y2)a𝐃(v)(ϕ,y2)\displaystyle\frac{\sigma_{N}^{4}}{N^{4}}\sum_{\begin{smallmatrix}Ns\leq Nu\leq Nt\\ Ns\leq Nv\leq Nt\end{smallmatrix}}\sum_{y^{1},y^{2}}\sum_{\mathbf{p}\in\mathbf{P}_{f}}\sum_{\begin{smallmatrix}\mathbf{A},\mathbf{B}\in\mathtt{F}(u)\\ \mathbf{C},\mathbf{D}\in\mathtt{F}(v)\\ \text{even intersections}\\ \mathbf{\iota}(\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D})=\mathbf{p}\end{smallmatrix}}\sigma_{N}^{|\mathbf{A}|+\dots+|\mathbf{D}|}a^{(u)}_{\mathbf{A}}(\phi,y^{1})a^{(u)}_{\mathbf{B}}(\phi,y^{1})a^{(v)}_{\mathbf{C}}(\phi,y^{2})a^{(v)}_{\mathbf{D}}(\phi,y^{2}) (4.5)

Next, we will see that the contributions to (4.2) from 𝒫2\mathcal{P}_{2} and 𝒫\mathcal{P}_{*} can be identified with the contribution from 𝒫1\mathcal{P}_{1}.

For fixed 0n<Nt0\leq n<Nt, we consider the contributions from 𝒫2\mathcal{P}_{2} and 𝒫\mathcal{P}_{*} to the summation in spatial variables at nn.

The contribution from 𝒫2\mathcal{P}_{2} at nn to (4.2) has the form of

Aa,Bb,Cc,DdAa+2,Bb+2,Cc+2,Dd+2Q(Aa,Bb,Cc,Dd)R(Aa+2,Bb+2,Cc+2,Dd+2)\displaystyle\sum_{\begin{smallmatrix}A_{a},B_{b},C_{c},D_{d}\\ A_{a+2},B_{b+2},C_{c+2},D_{d+2}\end{smallmatrix}}Q(A_{a},B_{b},C_{c},D_{d})R(A_{a+2},B_{b+2},C_{c+2},D_{d+2})
xyσN4[q(Aa,(n,x))q(Bb,(n,x))q(Cc,(n,y))q(Dd,(n,y))]\displaystyle\hskip 40.00006pt\cdot\sum_{x\not=y}\sigma_{N}^{4}\left[q(A_{a},(n,x))q(B_{b},(n,x))q(C_{c},(n,y))q(D_{d},(n,y))\right] (4.6)
[q((n,x),Aa+2)q((n,x),Bb+2)q((n,y),Cc+2)q((n,y),Dd+2)].\displaystyle\hskip 50.00008pt\cdot\left[q((n,x),A_{a+2})q((n,x),B_{b+2})q((n,y),C_{c+2})q((n,y),D_{d+2})\right].

Also, we can see from (3.2) that the contribution from 𝒫\mathcal{P}_{*} at nn to (4.2) has the form that replaces xyx\not=y by x=yx=y in the summation of (4.6).

Thus, we can identify the contribution from 𝒫\mathcal{P}_{*} at nn to (4.2) with the one from [AB][CD]𝒫2[AB][CD]\in\mathcal{P}_{2}. This is one-to-one correspondence.

Hence, we may consider that the summations in (4.5) of 𝐩\mathbf{p} are taken over the finite sequence in 𝒫1𝒫2\mathcal{P}_{1}\cup\mathcal{P}_{2}.

By a similar way, we can see that the contributions from [A][§]𝒫2[A\dagger][*\S]\in\mathcal{P}_{2} are identified with the one from A𝒫1A\dagger\in\mathcal{P}_{1}. This is one-to-one correspondence.

We write by 𝐣\mathbf{j} the map from 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} to 𝒫f={finite sequenice of AB,AC,AD,BC,BD,CD}\mathcal{P}_{f}=\{\text{finite sequenice of }AB,AC,AD,BC,BD,CD\} deduced from the above correspondence. We denote the length of 𝐩𝒫f\mathbf{p}\in\mathcal{P}_{f} by |𝐩||\mathbf{p}|.

Then, (4.5) can be rewritten by

σN4N4NsNuNtNsNvNty1,y2𝐩𝒫f𝐀,𝐁𝙵(u)𝐂,𝐃𝙵(v)even intersections𝐣(𝐀,𝐁,𝐂,𝐃)=𝐩σN2|𝐩|a𝐀(u)(ϕ,y1)a𝐁(u)(ϕ,y1)a𝐂(v)(ϕ,y2)a𝐃(v)(ϕ,y2)\displaystyle\frac{\sigma_{N}^{4}}{N^{4}}\sum_{\begin{smallmatrix}Ns\leq Nu\leq Nt\\ Ns\leq Nv\leq Nt\end{smallmatrix}}\sum_{y^{1},y^{2}}\sum_{\mathbf{p}\in\mathcal{P}_{f}}\sum_{\begin{smallmatrix}\mathbf{A},\mathbf{B}\in\mathtt{F}(u)\\ \mathbf{C},\mathbf{D}\in\mathtt{F}(v)\\ \text{even intersections}\\ \mathbf{j}(\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D})=\mathbf{p}\end{smallmatrix}}\sigma_{N}^{2|\mathbf{p}|}a^{(u)}_{\mathbf{A}}(\phi,y^{1})a^{(u)}_{\mathbf{B}}(\phi,y^{1})a^{(v)}_{\mathbf{C}}(\phi,y^{2})a^{(v)}_{\mathbf{D}}(\phi,y^{2}) (4.7)

For ϕ0\phi\geq 0, we can see that

σN4N4NsNuNtNsNvNty1,y2𝐩𝒫f𝐀,𝐁𝙵(u)𝐂,𝐃𝙵(v)even intersections𝐣(𝐀,𝐁,𝐂,𝐃)=𝐩σN2|𝐩|a𝐀(u)(ϕ,y1)a𝐁(u)(ϕ,y1)a𝐂(v)(ϕ,y2)a𝐃(v)(ϕ,y2)\displaystyle\frac{\sigma_{N}^{4}}{N^{4}}\sum_{\begin{smallmatrix}Ns\leq Nu\leq Nt\\ Ns\leq Nv\leq Nt\end{smallmatrix}}\sum_{y^{1},y^{2}}\sum_{\mathbf{p}\in\mathcal{P}_{f}}\sum_{\begin{smallmatrix}\mathbf{A},\mathbf{B}\in\mathtt{F}(u)\\ \mathbf{C},\mathbf{D}\in\mathtt{F}(v)\\ \text{even intersections}\\ \mathbf{j}(\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D})=\mathbf{p}\end{smallmatrix}}\sigma_{N}^{2|\mathbf{p}|}a^{(u)}_{\mathbf{A}}(\phi,y^{1})a^{(u)}_{\mathbf{B}}(\phi,y^{1})a^{(v)}_{\mathbf{C}}(\phi,y^{2})a^{(v)}_{\mathbf{D}}(\phi,y^{2})
σN4N4NsNuNtNsNvNty1,y2𝐩𝒫fx1,,x|𝐩|2(n1,,n|𝐩|)TN(𝐩,u,v)σN2|𝐩|a𝐀(u)(ϕ,y1)a𝐁(u)(ϕ,y1)a𝐂(v)(ϕ,y2)a𝐃(v)(ϕ,y2)\displaystyle\leq\frac{\sigma_{N}^{4}}{N^{4}}\sum_{\begin{smallmatrix}Ns\leq Nu\leq Nt\\ Ns\leq Nv\leq Nt\end{smallmatrix}}\sum_{y^{1},y^{2}}\sum_{\mathbf{p}\in\mathcal{P}_{f}}\sum_{x_{1},\dots,x_{|\mathbf{p}|}\in{\mathbb{Z}}^{2}}\sum_{(n_{1},\dots,n_{|\mathbf{p}|})\in T_{N}(\mathbf{p},u,v)}\sigma_{N}^{2|\mathbf{p}|}a^{(u)}_{\mathbf{A}}(\phi,y^{1})a^{(u)}_{\mathbf{B}}(\phi,y^{1})a^{(v)}_{\mathbf{C}}(\phi,y^{2})a^{(v)}_{\mathbf{D}}(\phi,y^{2}) (4.8)

where TN(𝐩,u,v)T_{N}(\mathbf{p},u,v) is the set of time-sequence {(n1,,n|𝐩|)}\{(n_{1},\dots,n_{|\mathbf{p}|})\} given as follows:

  1. (T-1)

    nin_{i} are associated with pip_{i} for 1i|𝐩|1\leq i\leq|\mathbf{p}|.

  2. (T-2)

    1n1n2n|𝐩|1\leq n_{1}\leq n_{2}\leq\dots\leq n_{|\mathbf{p}|}.

  3. (T-3)

    If pi=ABp_{i}=AB (or CDCD), then ni<Nun_{i}<Nu (ni<Nvn_{i}<Nv). Otherwise, ni<NuNvn_{i}<Nu\wedge Nv.

  4. (T-4)

    If pip_{i} and pi+1p_{i+1} is not a couple, then ni<ni+1n_{i}<n_{i+1}. Otherwise ni=ni+1n_{i}=n_{i+1} is allowed.

Remark 4.3.

𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D} does not appear in the sum on the right-hand side explicitly. However, 𝐩\mathbf{p} contains all their information.

Remark 4.4.

The inequality comes from the fact that ABCD𝒫ABCD\in\mathcal{P}_{*} is mapped to AB𝒫1AB\in\mathcal{P}_{1} so that the time-space summation associated with pAB,CDp\not=AB,CD does not contain the quadruple intersection. However, the difference is negligible. Indeed, the differences is dominated from above by the summation of quadruple intersection terms of expansion of ψ4E[𝖹N;tϕ(1)4]\|\psi\|_{\infty}^{4}E\left[\mathsf{Z}_{N;t}^{\phi}(1)^{4}\right]. However, we can find from the proof of [14, Theorem 6.1] that the contribution from the quadruple intersections is negligible (see the argument after Proposition 6.6 in [14]).

4.3 Partitions of sequence of pairings by stretches

Next, we focus on consecutive sequences in 𝐩𝒫f\mathbf{p}\in\mathcal{P}_{f}, called stretches in [12], that is, 𝐩=(p1,,pk)\mathbf{p}=(p_{1},\dots,p_{k}) can be divided into some blocks 𝚜=(s1=(p1,,p1),s2=(p1+1,,p2))\mathtt{s}=(s_{1}=(p_{1},\dots,p_{\ell_{1}}),s_{2}=(p_{\ell_{1}+1},\dots,p_{\ell_{2}})\dots), where we define

  1. (1)

    1=sup{j1:pjp1}\ell_{1}=\sup\{j\geq 1:p_{j}\not=p_{1}\}.

  2. (2)

    For each i1i\geq 1, i+1=sup{ji+1:pjpi+1}\ell_{i+1}=\sup\{j\geq\ell_{i}+1:p_{j}\not=p_{\ell_{i}+1}\} if i+1k\ell_{i}+1\leq k. Otherwise, we define i+1=\ell_{i+1}=\infty.

Thus, each block is associated with an element of P={AB,AC,AD,BC,BD,CD}P=\{AB,AC,AD,BC,BD,CD\}. We denote the number of stretches in 𝐩\mathbf{p} by |𝚜|=sup{k:k<}|\mathtt{s}|=\sup\{k:\ell_{k}<\infty\}.

We denote by 𝒫~f\widetilde{\mathcal{P}}_{f} the set of finite sequences 𝚜=(s1,s2,,s|𝚜|)\mathtt{s}=(s_{1},s_{2},\dots,s_{|\mathtt{s}|}) of 𝒫f\mathcal{P}_{f} with sisi+1s_{i}\not=s_{i+1} (i=1,,|𝚜|1i=1,\dots,|\mathtt{s}|-1).

We define the map 𝚔\mathtt{k} from 𝒫f\mathcal{P}_{f} to 𝒫~f\widetilde{\mathcal{P}}_{f} by 𝚔(𝐩)=𝚜\mathtt{k}(\mathbf{p})=\mathtt{s}.

Then, we can find that

σN4N4NsNuNtNsNvNty1,y2𝐩𝒫fx1,,x|𝐩|2(n1,,n|𝐩|)TN(𝐩,u,v)σN2|𝐩|a𝐀(u)(ϕ,y1)a𝐁(u)(ϕ,y1)a𝐂(v)(ϕ,y2)a𝐃(v)(ϕ,y2)\displaystyle\frac{\sigma_{N}^{4}}{N^{4}}\sum_{\begin{smallmatrix}Ns\leq Nu\leq Nt\\ Ns\leq Nv\leq Nt\end{smallmatrix}}\sum_{y^{1},y^{2}}\sum_{\mathbf{p}\in\mathcal{P}_{f}}\sum_{x_{1},\dots,x_{|\mathbf{p}|}\in{\mathbb{Z}}^{2}}\sum_{(n_{1},\dots,n_{|\mathbf{p}|})\in T_{N}(\mathbf{p},u,v)}\sigma_{N}^{2|\mathbf{p}|}a^{(u)}_{\mathbf{A}}(\phi,y^{1})a^{(u)}_{\mathbf{B}}(\phi,y^{1})a^{(v)}_{\mathbf{C}}(\phi,y^{2})a^{(v)}_{\mathbf{D}}(\phi,y^{2})
=σN4N4NsNuNtNsNvNty1,y2𝚜𝒫~f𝐩𝒫f𝚔(𝐩)=𝚜x1,,x|𝐩|2(n1,,n|𝐩|)TN(𝐩,u,v)σN2|𝐩|a𝐀(u)(ϕ,y1)a𝐁(u)(ϕ,y1)a𝐂(v)(ϕ,y2)a𝐃(v)(ϕ,y2)\displaystyle=\frac{\sigma_{N}^{4}}{N^{4}}\sum_{\begin{smallmatrix}Ns\leq Nu\leq Nt\\ Ns\leq Nv\leq Nt\end{smallmatrix}}\sum_{y^{1},y^{2}}\sum_{\mathtt{s}\in\widetilde{\mathcal{P}}_{f}}\sum_{\begin{smallmatrix}\mathbf{p}\in\mathcal{P}_{f}\\ \mathtt{k}(\mathbf{p})=\mathtt{s}\end{smallmatrix}}\sum_{x_{1},\dots,x_{|\mathbf{p}|}\in{\mathbb{Z}}^{2}}\sum_{(n_{1},\dots,n_{|\mathbf{p}|})\in T_{N}(\mathbf{p},u,v)}\sigma_{N}^{2|\mathbf{p}|}a^{(u)}_{\mathbf{A}}(\phi,y^{1})a^{(u)}_{\mathbf{B}}(\phi,y^{1})a^{(v)}_{\mathbf{C}}(\phi,y^{2})a^{(v)}_{\mathbf{D}}(\phi,y^{2})

Also, we can see that

𝚜𝒫~f𝐩𝒫f𝚔(𝐩)=𝚜x1,,x|𝐩|2(n1,,n|𝐩|)TN(𝐩,u,v)\displaystyle\sum_{\mathtt{s}\in\widetilde{\mathcal{P}}_{f}}\sum_{\begin{smallmatrix}\mathbf{p}\in\mathcal{P}_{f}\\ \mathtt{k}(\mathbf{p})=\mathtt{s}\end{smallmatrix}}\sum_{x_{1},\dots,x_{|\mathbf{p}|}\in{\mathbb{Z}}^{2}}\sum_{(n_{1},\dots,n_{|\mathbf{p}|})\in T_{N}(\mathbf{p},u,v)}\cdots
=𝚜𝒫~f(m1,x1),(n1,y1),,(m|𝚜|,x|𝚜|),(n|𝚜|,y|𝚜|)ST~N(𝚜,u,v)𝐩1,,𝐩|𝚜|𝒫f𝚔(𝐩i)=𝚜i(n1(i),x1(i)),,(n|𝐩i|2(i),x|𝐩i|2(i))ST^N(𝐩i,mi,ni),\displaystyle=\sum_{\mathtt{s}\in\widetilde{\mathcal{P}}_{f}}\sum_{(m_{1},x_{1}),(n_{1},y_{1}),\dots,(m_{|\mathtt{s}|},x_{|\mathtt{s}|}),(n_{|\mathtt{s}|},y_{|\mathtt{s}|})\in\widetilde{ST}_{N}(\mathtt{s},u,v)}\sum_{\begin{smallmatrix}\mathbf{p}_{1},\dots,\mathbf{p}_{|\mathtt{s}|}\in\mathcal{P}_{f}\\ \mathtt{k}(\mathbf{p}_{i})=\mathtt{s}_{i}\end{smallmatrix}}\sum_{(n^{(i)}_{1},x^{(i)}_{1}),\dots,(n^{(i)}_{|\mathbf{p}_{i}|-2},x^{(i)}_{|\mathbf{p}_{i}|-2})\in\widehat{ST}_{N}(\mathbf{p}_{i},m_{i},n_{i})}\cdots,

where ST~N(𝚜,u,v)\widetilde{ST}_{N}(\mathtt{s},u,v) is the set of space-time-sequence {(m1,x1),(n1,y1),,(m|𝚜|,x|𝚜|),(n|𝚜|,y|𝚜|)}\{(m_{1},x_{1}),(n_{1},y_{1}),\dots,(m_{|\mathtt{s}|},x_{|\mathtt{s}|}),(n_{|\mathtt{s}|},y_{|\mathtt{s}|})\} given as follows:

  1. (ST~\widetilde{\text{ST}}-1)

    (mi,xi),(ni,yi)(m_{i},x_{i}),(n_{i},y_{i}) are associated with the stretch sis_{i} for 1i|𝚜|1\leq i\leq|\mathtt{s}|, which represents the start point and the end point of the stretch.

  2. (ST~\widetilde{\text{ST}}-2)

    1m1n1m2n2m|𝚜|n|𝚜|1\leq m_{1}\leq n_{1}\leq m_{2}\leq n_{2}\leq\dots\leq m_{|\mathtt{s}|}\leq n_{|\mathtt{s}|}.

  3. (ST~\widetilde{\text{ST}}-3)

    If si=ABs_{i}=AB (CDCD), then ni<Nun_{i}<Nu (ni<Nvn_{i}<Nv). Otherwise, ni<NuNvn_{i}<Nu\wedge Nv.

  4. (ST~\widetilde{\text{ST}}-4)

    If sis_{i} and si+1s_{i+1} is not a couple, then ni<mi+1n_{i}<m_{i+1}. Otherwise ni=mi+1n_{i}=m_{i+1} is allowed.

  5. (ST~\widetilde{\text{ST}}-5)

    xi,yi2x_{i},y_{i}\in\mathbb{Z}^{2}.

Also, for 𝐩=(p1,,p|𝐩|)\mathbf{p}=(p_{1},\dots,p_{|\mathbf{p}|}), and 1mn1\leq m\leq n, ST^N(𝐩,m,n)\widehat{ST}_{N}(\mathbf{p},m,n) is the set of space-time-sequence {(n1,x1),,(n|𝐩|2,x|𝐩|2)}\{(n_{1},x_{1}),\dots,(n_{|\mathbf{p}|-2},x_{|\mathbf{p}|-2})\} given as follows:

  1. (ST^\widehat{\text{ST}}-1)

    (ni,xi)(n_{i},x_{i}) are associated with pip_{i} for 1i|𝐩|21\leq i\leq|\mathbf{p}|-2.

  2. (ST^\widehat{\text{ST}}-2)

    m<n1<n2<<n|𝐩|2<nm<n_{1}<n_{2}<\dots<n_{|\mathbf{p}|-2}<n.

  3. (ST^\widehat{\text{ST}}-3)

    xi2x_{i}\in\mathbb{Z}^{2}.

Now, we focus on the summation over ST^N(𝐩i,mi,ni)\widehat{ST}_{N}(\mathbf{p}_{i},m_{i},n_{i}), and 𝐩i\mathbf{p}_{i}. Fix (mi,xi)(m_{i},x_{i}) and (ni,yi)(n_{i},y_{i}). Then, the other variables appear in the summand with the form

{V(X)σN2|𝐩i|1xi if |𝐩i|=1V(X)σN2|𝐩i|qnimi(xi,yi)2 if |𝐩i|=2V(X)σN2|𝐩i|qn(i)1mi(xi,x(i)1)2j=1|𝐩|2qn(i)(j+1)n(i)(j)(x(i)j+1,x(i)j)2qnin(i)|𝐩i|1(x(i)|𝐩i|1,yi)2 if |𝐩i|3,\displaystyle\begin{cases}V(X)\sigma_{N}^{2|\mathbf{p}_{i}|}1_{x_{i}}&\text{ if }|\mathbf{p}_{i}|=1\\ V(X)\sigma_{N}^{2|\mathbf{p}_{i}|}q_{n_{i}-m_{i}}(x_{i},y_{i})^{2}&\text{ if }|\mathbf{p}_{i}|=2\\ V(X)\sigma_{N}^{2|\mathbf{p}_{i}|}q_{n_{(i)}^{1}-m_{i}}(x_{i},x_{(i)}^{1})^{2}\prod_{j=1}^{|\mathbf{p}|-2}q_{n_{(i)}^{(j+1)}-n_{(i)}^{(j)}}(x_{(i)}^{j+1},x_{(i)}^{j})^{2}q_{n_{i}-n_{(i)}^{|\mathbf{p}_{i}|-1}}(x_{(i)}^{|\mathbf{p}_{i}|-1},y_{i})^{2}&\text{ if }|\mathbf{p}_{i}|\geq 3,\end{cases}

where V(X)V(X) is a function independent of the summation. Hence, it has the given by

V(X)Umi,ni(xi,yi)\displaystyle V(X)U_{m_{i},n_{i}}(x_{i},y_{i})

Repeating this procedure, we rewrite (4.8) by the following form:

σN4N4NsNuNtNsNvNtz1,z2𝚜𝒫~f(m1,x1),(n1,y1),,(m|𝚜|,x|𝚜|),(n|𝚜|,y|𝚜|)ST~N(𝚜,u,v)FN(ϕ,ψ,𝐦,𝐧,𝐱,𝐲)i=1|𝚜|(Umi,ni(xi,yi)).\displaystyle\frac{\sigma_{N}^{4}}{N^{4}}\sum_{\begin{smallmatrix}Ns\leq Nu\leq Nt\\ Ns\leq Nv\leq Nt\end{smallmatrix}}\sum_{z_{1},z_{2}}\sum_{\mathtt{s}\in\widetilde{\mathcal{P}}_{f}}\sum_{(m_{1},x_{1}),(n_{1},y_{1}),\dots,(m_{|\mathtt{s}|},x_{|\mathtt{s}|}),(n_{|\mathtt{s}|},y_{|\mathtt{s}|})\in\widetilde{ST}_{N}(\mathtt{s},u,v)}F_{N}(\phi,\psi,\mathbf{m},\mathbf{n},\mathbf{x},\mathbf{y})\prod_{i=1}^{|\mathtt{s}|}\left(U_{m_{i},n_{i}}(x_{i},y_{i})\right).

To give the explicit form of FN(ϕ,ψ,𝐦,𝐧,𝐱,𝐲)F_{N}(\phi,\psi,\mathbf{m},\mathbf{n},\mathbf{x},\mathbf{y}), we will see the sequence of (mi,xi)(m_{i},x_{i}), (ni,yi)(n_{i},y_{i}).

Definition 4.5.

For each 𝚜𝒫~f\mathtt{s}\in\widetilde{\mathcal{P}}_{f} and E{A,B,C,D}E\in\{A,B,C,D\}, we set

i1E=inf{j1:sjE},\displaystyle i^{E}_{1}=\inf\{j\geq 1:s_{j}\ni E\},
and if ikE<,ik+1E=inf{j>ikE:sjE},\displaystyle\text{and if $i^{E}_{k}<\infty$},i_{k+1}^{E}=\inf\{j>i_{k}^{E}:s_{j}\ni E\},

where we set inf=\inf=\infty. Also, we denote by kE=sup{k:ik+1E=}k^{E}=\sup\{k:i_{k+1}^{E}=\infty\} the number of times EE appears in 𝚜\mathtt{s}.

Also, we set

mjE=mijE,njE=nijE,xjE=xijE, and yjE=yijE\displaystyle m^{E}_{j}=m_{i^{E}_{j}},n^{E}_{j}=n_{i^{E}_{j}},x^{E}_{j}=x_{i^{E}_{j}},\text{ and }y^{E}_{j}=y_{i^{E}_{j}}

for m1,n1,,mk,nkm_{1},n_{1},\dots,m_{k},n_{k}, x1,y1,,xk,yk2x_{1},y_{1},\dots,x_{k},y_{k}\in{\mathbb{Z}}^{2}, and 𝚜𝒫~f\mathtt{s}\in\widetilde{\mathcal{P}}_{f}, j=1,,kEj=1,\dots,k^{E} and E{A,B,C,D}E\in\{A,B,C,D\}.

For each E{A,B,C,D}E\in\{A,B,C,D\}, the transition between sijEs_{i^{E}_{j}} and sij+1Es_{i^{E}_{j+1}} is from (njE,yjE)(n^{E}_{j},y^{E}_{j}) to (mj+1E,xj+1E)(m^{E}_{j+1},x^{E}_{j+1}). Its contributions to FN(ϕ,ψ,𝐦,𝐧,𝐱,𝐲)F_{N}(\phi,\psi,\mathbf{m},\mathbf{n},\mathbf{x},\mathbf{y}) are given by the form GNqmj+1EnjE(yjE,xj+1E)G_{N}\cdot q_{m^{E}_{j+1}-n^{E}_{j}}\left(y^{E}_{j},x^{E}_{j+1}\right) for some GNG_{N}. Thus, we can see that

(4.8)\displaystyle\eqref{eq:partmoment33} =σN4N4NsNuNtNsNvNtz1,z2𝚜𝒫~f(m1,x1),(n1,y1),,(m|𝚜|,x|𝚜|),(n|𝚜|,y|𝚜|)ST~N(𝚜,u,v)\displaystyle=\frac{\sigma_{N}^{4}}{N^{4}}\sum_{\begin{smallmatrix}Ns\leq Nu\leq Nt\\ Ns\leq Nv\leq Nt\end{smallmatrix}}\sum_{z_{1},z_{2}}\sum_{\mathtt{s}\in\widetilde{\mathcal{P}}_{f}}\sum_{(m_{1},x_{1}),(n_{1},y_{1}),\dots,(m_{|\mathtt{s}|},x_{|\mathtt{s}|}),(n_{|\mathtt{s}|},y_{|\mathtt{s}|})\in\widetilde{ST}_{N}(\mathtt{s},u,v)}
E{A,B,C,D}(qm1E(ϕN,x1E)j=1kE1qmj+1EnjE(yjE,xj+1E))i=1|𝚜|(Umi,ni(xi,yi))\displaystyle\prod_{E\in\{A,B,C,D\}}\left(q_{m_{1}^{E}}(\phi_{N},x_{1}^{E})\prod_{j=1}^{k_{E}-1}q_{m^{E}_{j+1}-n^{E}_{j}}\left(y^{E}_{j},x^{E}_{j+1}\right)\right)\prod_{i=1}^{|\mathtt{s}|}\left(U_{m_{i},n_{i}}(x_{i},y_{i})\right)
qNunkA(ykA,z1)qNunkB(ykB,z1)qNvnkC(ykC,z2)qNvnkD(ykD,z2)ψN(z1)2ψN(z2)2.\displaystyle q_{Nu-n_{k^{A}}}(y_{k^{A}},z_{1})q_{Nu-n_{k^{B}}}(y_{k^{B}},z_{1})q_{Nv-n_{k^{C}}}(y_{k^{C}},z_{2})q_{Nv-n_{k^{D}}}(y_{k^{D}},z_{2})\psi_{N}(z_{1})^{2}\psi_{N}(z_{2})^{2}.

Moreover, we remark that the summation

σN4NsNuNtNsNvNtz1,z2\displaystyle\sigma_{N}^{4}\sum_{\begin{smallmatrix}Ns\leq Nu\leq Nt\\ Ns\leq Nv\leq Nt\end{smallmatrix}}\sum_{z_{1},z_{2}}\cdots

can be embedded into the summation of ABAB and CDCD. Hence, we have

(4.8)\displaystyle\eqref{eq:partmoment33} =2N4𝚜𝒫~f,AB,CDx1,,x|𝚜|y1,,y|𝚜|in2(m1,n1,,(m|𝚜|,n|𝚜|)T~N(𝚜,s,t)\displaystyle=\frac{2}{N^{4}}\sum_{\mathtt{s}\in\widetilde{\mathcal{P}}_{f,AB,CD}}\sum_{\begin{smallmatrix}x_{1},\dots,x_{|\mathtt{s}|}\\ y_{1},\dots,y_{|\mathtt{s}|}in{\mathbb{Z}}^{2}\end{smallmatrix}}\sum_{(m_{1},n_{1},\dots,(m_{|\mathtt{s}|},n_{|\mathtt{s}|})\in\widetilde{T}_{N}(\mathtt{s},s,t)}
E{A,B,C,D}(qm1E(ϕN,x1E)j=1kE1qmj+1EnjE(yjE,xj+1E))i=1|𝚜|(Umi,ni(xi,yi))ψN(y|𝚜|1)2ψN(y|𝚜|)2,\displaystyle\prod_{E\in\{A,B,C,D\}}\left(q_{m_{1}^{E}}(\phi_{N},x_{1}^{E})\prod_{j=1}^{k_{E}-1}q_{m^{E}_{j+1}-n^{E}_{j}}\left(y^{E}_{j},x^{E}_{j+1}\right)\right)\prod_{i=1}^{|\mathtt{s}|}\left(U_{m_{i},n_{i}}(x_{i},y_{i})\right)\psi_{N}(y_{|\mathtt{s}|-1})^{2}\psi_{N}(y_{|\mathtt{s}|})^{2}, (4.9)

where we set

𝒫~f,AB,CD={𝚜=(s1,,sk)P:sk1=AB,sk=CD,k2}\displaystyle\widetilde{\mathcal{P}}_{f,AB,CD}=\{\mathtt{s}=(s_{1},\dots,s_{k})\in P:s_{k-1}=AB,s_{k}=CD,k\geq 2\}

and T~N(𝚜,s,t)\widetilde{T}_{N}(\mathtt{s},s,t) is the set of sequence of time-pairs {(m1,n1),,(m|𝚜|,n|𝚜|)}\{(m_{1},n_{1}),\dots,(m_{|\mathtt{s}|},n_{|\mathtt{s}|})\} satisfy the followings:

  1. (T~\widetilde{\text{T}}-1)

    (mi,ni)(m_{i},n_{i}) are associated with the stretch sis_{i} for 1i|𝚜|1\leq i\leq|\mathtt{s}|, which represents the start time and the end time of the stretch.

  2. (T~\widetilde{\text{T}}-2)

    1m1n1m2n2m|𝚜|n|𝚜|1\leq m_{1}\leq n_{1}\leq m_{2}\leq n_{2}\leq\dots\leq m_{|\mathtt{s}|}\leq n_{|\mathtt{s}|}.

  3. (T~\widetilde{\text{T}}-3)

    Nsn|𝚜|1m|𝚜|n|𝚜|<NtNs\leq n_{|\mathtt{s}|-1}\leq m_{|\mathtt{s}|}\leq n_{|\mathtt{s}|}<Nt.

  4. (T~\widetilde{\text{T}}-4)

    If sis_{i} and si+1s_{i+1} are not a couple, then ni<mi+1n_{i}<m_{i+1}. Otherwise, ni=mi+1n_{i}=m_{i+1} is allowed.

In particular, the summand is given as the products of the weights associated with the graphs.

Thus, it is enough to estimate (4.9).

We now introduce a new oriented graph G(𝚜)=(V(𝚜),E(𝚜))G(\mathtt{s})=(V(\mathtt{s}),E(\mathtt{s})) with vertices V(𝚜)={0,1,,|𝚜|}V(\mathtt{s})=\{0,1,\dots,|\mathtt{s}|\}, where the oriented edges are [0,i1E\left[0,i_{1}^{E}\right\rangle (E{A,B,C,D}E\in\{A,B,C,D\}) and [ijE,ij+1E)\left[i_{j}^{E},i_{j+1}^{E}\right) for 1jkE11\leq j\leq k^{E}-1 (E{A,B,C,D}E\in\{A,B,C,D\}).

We write

q(e)={qmi1E(ϕN,xi1E)for e=[0,i1E) (E{A,B,C,D}qmij+1EnijE(yijE,xij+1E)for e=[ijE,ij+1E) for 1jkE1 (E{A,B,C,D})\displaystyle q(e)=\begin{cases}q_{m_{i_{1}}^{E}}(\phi_{N},x_{i_{1}^{E}})\quad&\text{for $e=\left[0,i_{1}^{E}\right)$ ($E\in\{A,B,C,D\}$) }\\ q_{m_{i_{j}+1}^{E}-n_{i_{j}}^{E}}(y_{i_{j}^{E}},x_{i_{j+1}^{E}})\quad&\text{for $e=\left[i_{j}^{E},i_{j+1}^{E}\right)$ for $1\leq j\leq k^{E}-1$ ($E\in\{A,B,C,D\}$)}\end{cases}
and
U(v)=Umv,nv(xv,yv)for v{1,,|𝚜|}.\displaystyle U(v)=U_{m_{v},n_{v}}(x_{v},y_{v})\quad\text{for $v\in\{1,\dots,|\mathtt{s}|\}$}.

Then,

(4.9)=2N4𝚜𝒫~f,AB,CDx1,,x|𝚜|y1,,y|𝚜|in2{(m1,n1),,(m|𝚜|,n|𝚜|)}T~N(𝚜,s,t)eE(𝚜)q(e)i=1|𝚜|U(i).\displaystyle\eqref{eq:partmoment34}=\frac{2}{N^{4}}\sum_{\mathtt{s}\in\widetilde{\mathcal{P}}_{f,AB,CD}}\sum_{\begin{smallmatrix}x_{1},\dots,x_{|\mathtt{s}|}\\ y_{1},\dots,y_{|\mathtt{s}|}in{\mathbb{Z}}^{2}\end{smallmatrix}}\sum_{\{(m_{1},n_{1}),\dots,(m_{|\mathtt{s}|},n_{|\mathtt{s}|})\}\in\widetilde{T}_{N}(\mathtt{s},s,t)}\prod_{e\in E(\mathtt{s})}q(e)\prod_{i=1}^{|\mathtt{s}|}U(i). (4.10)
Refer to caption
Figure 1: An image of the graph associated with 𝚜\mathtt{s} and T~N(𝚜,s,t)\widetilde{T}_{N}(\mathtt{s},s,t). Curly lines represent wights UU and solid lines represent weights qq.

We can see the following structure of the graph G(𝚜)G(\mathtt{s}).

Definition 4.6.

Each v{1,,|𝚜|2}v\in\{1,\dots,|\mathtt{s}|-2\} has two incoming edges [Iv,v[I_{v},v\rangle , [Iv,v[I_{v}^{\prime},v\rangle and two outgoing edge [v,Ov[v,O_{v}\rangle and [v,Ov[v,O_{v}^{\prime}\rangle, where 0IvIv0\leq I_{v}\leq I_{v}^{\prime} and OvOvO_{v}^{\prime}\leq O_{v}.

Let l(𝚜)=sup{i1E:E=A,B,C,D}l(\mathtt{s})=\sup\left\{i_{1}^{E}:E=A,B,C,D\right\}.

For simplicity, we set i1A=i1B=1i_{1}^{A}=i_{1}^{B}=1, i1C=2i_{1}^{C}=2, and i1D2i_{1}^{D}\geq 2. (The other cases are obtained by permutation.)

Proposition 4.7.

For each 𝚜\mathtt{s}, the following holds.

  1. (1)

    I1=I1=0I_{1}=I^{\prime}_{1}=0. Also, the following holds:

    1. (i)

      If s1s_{1} and s2s_{2} are a couple, then i1D=2i_{1}^{D}=2 and I2=I2=0I_{2}=I_{2}^{\prime}=0.

    2. (ii)

      If s1s_{1} and s2s_{2} are not a couple, then i1D3i_{1}^{D}\geq 3, I2=0I_{2}=0, I2=1I_{2}^{\prime}=1, Ii1D=0I_{i_{1}^{D}}=0, and Ii1D{i1D2,i1D1}I_{i_{1}^{D}}^{\prime}\in\{i_{1}^{D}-2,i_{1}^{D}-1\}. In particular, Ii1D=i1D2I_{i_{1}^{D}}^{\prime}=i_{1}^{D}-2 if and only if si1Ds_{i_{1}^{D}} and si1Ds_{i_{1}^{D}} are a couple.

  2. (2)

    Let i1,2,i1Di\not=1,2,i_{1}^{D}.

    1. (i)

      If si1s_{i-1} and sis_{i} are a couple, then Ii=i2IiI_{i}^{\prime}=i-2\geq I_{i} and the equality holds if and only if si2=sis_{i-2}=s_{i}.

    2. (ii)

      If si1s_{i-1} and sis_{i} are not a couple, then Ii=i1I_{i}^{\prime}=i-1 and Iii2I_{i}\leq i-2. In particular, Ii=i2I_{i}=i-2 if and only if each label in sis_{i} is contained in either si2s_{i-2} or si1s_{i-1}.

  3. (3)

    For 1il(𝚜)21\leq i\leq l(\mathtt{s})-2, Oi=i+1O_{i}^{\prime}=i+1, Oi=i+2O_{i}=i+2.

  4. (4)

    Let l(𝚜)2i|𝚜|2l(\mathtt{s})-2\leq i\leq|\mathtt{s}|-2. If sis_{i} and si+1s_{i+1} are a couple, then Oi=i+2OiO_{i}^{\prime}=i+2\leq O_{i}. If sis_{i} and si+1s_{i+1} are not a couple, Oi=i+1O_{i}^{\prime}=i+1 and Oii+2O_{i}\geq i+2. In particular, Oi=i+2O_{i}=i+2 if and only if each label in sis_{i} are contained in si+1s_{i+1} or si+2s_{i+2}

Proof.

(1) I1=I1=0I_{1}=I_{1}^{\prime}=0 is trivial by definition.

(1)(i) If s1=ABs_{1}=AB and s2s_{2} are a couple, then s2=CDs_{2}=CD so that i1C=i1D=2i_{1}^{C}=i_{1}^{D}=2 and hence I2=I2=0I_{2}=I_{2}^{\prime}=0.

(1)(ii) If s1=ABs_{1}=AB and s2s_{2} are not a couple, then s2=Cs_{2}=*C ({A,B}*\in\{A,B\}). So i1D3i_{1}^{D}\geq 3 and there exists oriented edges [1,2[1,2\rangle and [0,2[0,2\rangle. Also, it is trivial that Ii1D=0I_{i_{1}^{D}}=0. Finally, if si1D1s_{i_{1}^{D}-1} and si1Ds_{i_{1}^{D}} are a couple (e.g. si1D1=ABs_{i_{1}^{D}-1}=AB and si1D=CDs_{i_{1}^{D}}=CD), then si1D2=ECs_{i_{1}^{D}-2}=EC for E{A,B}E\in\{A,B\} since it does not contain DD and si1D2si1D1s_{i_{1}^{D}-2}\not=s_{i_{1}^{D}-1}. Thus, Ii1D=i2I_{i_{1}^{D}}^{\prime}=i-2. On the other hand, if si1D1s_{i_{1}^{D}-1} and si1Ds_{i_{1}^{D}} are not a couple, then Ii1D=i1I_{i_{1}^{D}}^{\prime}=i-1 holds by definition.

(2)

(2)(i) If si1s_{i-1} and sis_{i} are a couple (e.g. si1=ABs_{i-1}=AB and si=CDs_{i}=CD), then si2=s_{i-2}=*\dagger for {A,B,C,D}*\in\{A,B,C,D\} and {C,D}\dagger\in\{C,D\} since si1si2s_{i-1}\not=s_{i-2}. Therefore, Ii=i2I_{i}^{\prime}=i-2. Also, Ii=i2I_{i}=i-2 if and only if there exist k,lk,l such that ikC=ilD=i2i_{k}^{C}=i_{l}^{D}=i-2 and ik+1C=ij+1D=ii_{k+1}^{C}=i_{j+1}^{D}=i, i.e. si2=sis_{i-2}=s_{i}.

(2)(ii) If si1s_{i-1} and sis_{i} are not a couple (e.g. si1=ABs_{i-1}=AB and si=BCs_{i}=BC), then Ii=i1I_{i}^{\prime}=i-1 and Iii2I_{i}\leq i-2. If si2=Ds_{i-2}=*D for {A,B,C}*\in\{A,B,C\}, then [i2,i[i-2,i\rangle exists. On the other hand, if si2=s_{i-2}=*\dagger for ,{A,B,C}*,\dagger\in\{A,B,C\}, then [i2,i[i-2,i\rangle does not exist so Ii<i2I_{i}<i-2.

(3) By definition, the labels contained in s1,,sl(𝚜)1s_{1},\dots,s_{l(\mathtt{s})-1} are A,B,CA,B,C. Then, for 1il(𝚜)21\leq i\leq l(\mathtt{s})-2, sis_{i} and si+1s_{i+1} are not a pair(e.g. si=ABs_{i}=AB and si+1=ACs_{i+1}=AC), and si+1s_{i+1} and si+2s_{i+2} are not a pair and hence si+2=ABs_{i+2}=AB or BCBC.

(4) The proof is the same as (2).

Now, we will give an upper bound of (4.9) by taking summation in spatial variables xi,yix_{i},y_{i}.

First, we will take summation in the order of y|𝚜|y_{|\mathtt{s}|}\rightarrowx|𝚜|x_{|\mathtt{s}|}\rightarrowy|𝚜|1y_{|\mathtt{s}|-1}\rightarrow\dots as follows: We remark that xix_{i} (1i|𝚜|1\leq i\leq|\mathtt{s}|) appear in U(i)U(i) just one time and in q(e1)q(e_{1}) and q(e2)q(e_{2}) for just two e1,e2e_{1},e_{2} and the same holds for yiy_{i} (1i|𝚜|21\leq i\leq|\mathtt{s}|-2).

The summand of (4.10) has the form F(𝐦,𝐧,x1,,x|𝚜|,y1,,y|𝚜|1)Um|𝚜|,n|𝚜|(x|𝚜|,y|𝚜|)F(\mathbf{m},\mathbf{n},x_{1},\dots,x_{|\mathtt{s}|},y_{1},\dots,y_{|\mathtt{s}|-1})U_{m_{|\mathtt{s}|},n_{|\mathtt{s}|}}(x_{|\mathtt{s}|},y_{|\mathtt{s}|}), and hence the summation of (4.10) in y|𝚜|y_{|\mathtt{s}|}) is dominated by F(𝐦,𝐧,x1,,x|𝚜|,y1,,y|𝚜|1)Um|𝚜|,n|𝚜|F(\mathbf{m},\mathbf{n},x_{1},\dots,x_{|\mathtt{s}|},y_{1},\dots,y_{|\mathtt{s}|-1})U_{m_{|\mathtt{s}|},n_{|\mathtt{s}|}}. In particular, x|𝚜|x_{|\mathtt{s}|} appears as q([I|𝚜|,|𝚜|)q([I|𝚜|,|𝚜|)q([I_{|\mathtt{s}|},|\mathtt{s}|\rangle)q([I_{|\mathtt{s}|}^{\prime},|\mathtt{s}|\rangle) in F(𝐦,𝐧,x1,,x|𝚜|,y1,,y|𝚜|1)F(\mathbf{m},\mathbf{n},x_{1},\dots,x_{|\mathtt{s}|},y_{1},\dots,y_{|\mathtt{s}|-1}).

We know that for j1j\geq 1,

xjq([Ij,j)q([Ij,j)\displaystyle\sum_{x_{j}}q([I_{j},j\rangle)q([I_{j}^{\prime},j\rangle) =q2mjnIjnIj(yIj,Ij)\displaystyle=q_{2m_{j}-n_{I_{j}}-n_{I_{j}^{\prime}}}(y_{I_{j}},I_{j}^{\prime})
Cq,12mjnIjnIj\displaystyle\leq\frac{C_{q,1}}{2m_{j}-n_{I_{j}}-n_{I_{j}^{\prime}}} (4.11)

if 1IjIj1\leq I_{j}\leq I_{j}^{\prime}, where Cq,1C_{q,1} is a constant which is uniformly chosen in mi,nIi,nIim_{i},n_{I_{i}},n_{I_{i}^{\prime}} and yIj,yIjy_{I_{j}},y_{I_{j}^{\prime}},

xjq([0,j)q([Ij,j)=q2mjnIj(ϕN,yIj)Cq,2\displaystyle\sum_{x_{j}}q([0,j\rangle)q([I_{j}^{\prime},j\rangle)=q_{2m_{j}-n_{I_{j}^{\prime}}}(\phi_{N},y_{I_{j}}^{\prime})\leq C_{q,2} (4.12)

if Ij=0<IjI_{j}=0<I_{j}^{\prime}, where Cq,2C_{q,2} is a constant depending only on ϕ(x)dx\int\phi(x)\text{\rm d}x, and

xjq([Ij,j)q([0,j)\displaystyle\sum_{x_{j}}q([I_{j},j\rangle)q([0,j\rangle) =yeven2q2mj(ϕN,y)ϕN(y)Cq,3N\displaystyle=\sum_{y\in{\mathbb{Z}}^{2}_{\mathrm{even}}}q_{2m_{j}}(\phi_{N},y)\phi_{N}(y)\leq C_{q,3}N (4.13)

if Ij=Ij=0I_{j}=I_{j}=0, where Cq,3C_{q,3} is a constant depending only on ϕ(x)dx\int\phi(x)\text{\rm d}x. We denote by

q~(j)={Cq,12mjnIjnIjif 1IjIjCq,2if 0=Ij<IjCq,3Nif Ij=Ij=0.\displaystyle\widetilde{q}(j)=\begin{cases}\frac{C_{q,1}}{2m_{j}-n_{I_{j}}-n_{I_{j}^{\prime}}}\quad&\text{if }1\leq I_{j}\leq I_{j}^{\prime}\\ C_{q,2}&\text{if }0=I_{j}<I_{j}^{\prime}\\ C_{q,3}N&\text{if }I_{j}=I_{j}^{\prime}=0.\end{cases}

Hence, yI|𝚜|y_{I_{|\mathtt{s}|}} and yI|𝚜|y_{I_{|\mathtt{s}|}^{\prime}} appear in the summand with the form U(I𝚜)q([Iv,OIv)U(I|𝚜|)q([I|𝚜|,OI|𝚜|)U(I_{\mathtt{s}})q\left(\left[I_{v},O^{\prime}_{I_{v}}\right\rangle\right)U(I_{|\mathtt{s}|}^{\prime})q([I_{|\mathtt{s}|}^{\prime},O^{\prime}_{I_{|\mathtt{s}|}^{\prime}}\rangle) (if I|𝚜|<I|𝚜|I_{|\mathtt{s}|}<I_{|\mathtt{s}|}^{\prime}) or U(I|𝚜|)U(I_{|\mathtt{s}|}) (if I|𝚜|=I|𝚜|I_{{|\mathtt{s}|}}=I^{\prime}_{|\mathtt{s}|}).

Let 1j|𝚜|11\leq j\leq|\mathtt{s}|-1. Suppose that by taking summation in xj+1,,x|𝚜|x_{j+1},\dots,x_{|\mathtt{s}|} and yj+1,,y|𝚜|y_{j+1},\dots,y_{|\mathtt{s}|}, the summand has the form

VN(𝐦,𝐧,j)eEO(𝚜,j)q(e)i=1jU(i),\displaystyle V_{N}(\mathbf{m},\mathbf{n},j)\prod_{e\in E_{O}(\mathtt{s},j)}q(e)\prod_{i=1}^{j}U(i), (4.14)
EO(𝚜,j)={eE(𝚜):e=[i,Oi (Oij)e=[i,Oi (Oij)e=[0,i1E (i1Ej)}.\displaystyle E_{O}(\mathtt{s},j)=\left\{e\in E(\mathtt{s}):\begin{tabular}[]{l}$e=[i,O_{i}\rangle$ \quad$(O_{i}\leq j)$\\ $e=[i,O^{\prime}_{i}\rangle$ \quad$(O^{\prime}_{i}\leq j)$\\ $e=[0,i_{1}^{E}\rangle$ \quad$(i_{1}^{E}\leq j)$\end{tabular}\right\}. (4.18)

Since yjy_{j} appears only in U(i)U(i), the summation in yjy_{j} of (4.14) is dominated by

Umj,njVN(𝐦,𝐧,j)eEO(𝚜,j)q(e)i=1j1U(i)\displaystyle U_{m_{j},n_{j}}V_{N}(\mathbf{m},\mathbf{n},j)\prod_{e\in E_{O}(\mathtt{s},j)}q(e)\prod_{i=1}^{j-1}U(i) (4.19)

and xjx_{j} appears as q([Ij,j)q([Ij,j)q([I_{j},j\rangle)q([I_{j}^{\prime},j\rangle) in (4.19). The summation of (4.19) in xjx_{j} is dominated by

q~(j)Umj,njVN(𝐦,𝐧,j)eEO(𝚜,j1)q(e)i=1j1U(i).\displaystyle\widetilde{q}(j)U_{m_{j},n_{j}}V_{N}(\mathbf{m},\mathbf{n},j)\prod_{e\in E_{O}(\mathtt{s},j-1)}q(e)\prod_{i=1}^{j-1}U(i).

By induction, we can obtain an upper bound of (4.10). To give it, we divide 𝒫~f,AB,CD\widetilde{\mathcal{P}}_{f,AB,CD} into two disjoint sets

𝒫~α={𝚜𝒫~f,AB,CD:(s1,s2) are not a couple.}\displaystyle\widetilde{\mathcal{P}}_{\alpha}=\{\mathtt{s}\in\widetilde{\mathcal{P}}_{f,AB,CD}:\text{$(s_{1},s_{2})$ are not a couple.}\}
𝒫~β={𝚜𝒫~f,AB,CD:(s1,s2) are a couple.}.\displaystyle\widetilde{\mathcal{P}}_{\beta}=\{\mathtt{s}\in\widetilde{\mathcal{P}}_{f,AB,CD}:(s_{1},s_{2})\text{ are a couple.}\}.

For 𝚜𝒫~α\mathtt{s}\in\widetilde{\mathcal{P}}_{\alpha}, one (4.13) and two (4.12) appear. On the other hand, for 𝚜𝒫~β\mathtt{s}\in\widetilde{\mathcal{P}}_{\beta}, two (4.13) and no (4.12) appear.

Thus, we can find that (4.10) is dominated by

2Cq,2Cq,3N3𝚜𝒫~α(𝐦,𝐧)T~N(𝚜,s,t)j=3,,|𝚜|jl(𝚜)q~(j)j=1|𝚜|Umj,nj\displaystyle\frac{2C_{q,2}C_{q,3}}{N^{3}}\sum_{\mathtt{s}\in\widetilde{\mathcal{P}}_{\alpha}}\sum_{(\mathbf{m},\mathbf{n})\in\widetilde{T}_{N}(\mathtt{s},s,t)}\prod_{\begin{smallmatrix}j=3,\dots,|\mathtt{s}|\\ j\not=l(\mathtt{s})\end{smallmatrix}}\widetilde{q}(j)\prod_{j=1}^{|\mathtt{s}|}U_{m_{j},n_{j}}
+2Cq,32N2𝚜𝒫~β(𝐦,𝐧)T~N(𝚜,s,t)j=3,,|𝚜|q~(j)j=1|𝚜|Umj,nj\displaystyle+\frac{2C_{q,3}^{2}}{N^{2}}\sum_{\mathtt{s}\in\widetilde{\mathcal{P}}_{\beta}}\sum_{(\mathbf{m},\mathbf{n})\in\widetilde{T}_{N}(\mathtt{s},s,t)}\prod_{\begin{smallmatrix}j=3,\dots,|\mathtt{s}|\end{smallmatrix}}\widetilde{q}(j)\prod_{j=1}^{|\mathtt{s}|}U_{m_{j},n_{j}}
2Cq,2Cq,3e2TλN3𝚜𝒫~α(𝐦,𝐧)T~N(𝚜,s,t)j=3,,|𝚜|jl(𝚜)q~(j)j=1|𝚜|eλnjmjNUmj,nj\displaystyle\leq\frac{2C_{q,2}C_{q,3}e^{2T\lambda}}{N^{3}}\sum_{\mathtt{s}\in\widetilde{\mathcal{P}}_{\alpha}}\sum_{(\mathbf{m},\mathbf{n})\in\widetilde{T}_{N}(\mathtt{s},s,t)}\prod_{\begin{smallmatrix}j=3,\dots,|\mathtt{s}|\\ j\not=l(\mathtt{s})\end{smallmatrix}}\widetilde{q}(j)\prod_{j=1}^{|\mathtt{s}|}e^{-\lambda\frac{n_{j}-m_{j}}{N}}U_{m_{j},n_{j}} (Type-α\alpha)
+2Cq,32e2TλN2𝚜𝒫~β(𝐦,𝐧)T~N(𝚜,s,t)j=3,,|𝚜|q~(j)j=1|𝚜|eλnjmjNUmj,nj.\displaystyle+\frac{2C_{q,3}^{2}e^{2T\lambda}}{N^{2}}\sum_{\mathtt{s}\in\widetilde{\mathcal{P}}_{\beta}}\sum_{(\mathbf{m},\mathbf{n})\in\widetilde{T}_{N}(\mathtt{s},s,t)}\prod_{\begin{smallmatrix}j=3,\dots,|\mathtt{s}|\end{smallmatrix}}\widetilde{q}(j)\prod_{j=1}^{|\mathtt{s}|}e^{-\lambda\frac{n_{j}-m_{j}}{N}}U_{m_{j},n_{j}}. (Type-β\beta)

Here λ>0\lambda>0 is a constant (chosen later) which will play the same role as the one introduced in [12].

[Uncaptioned image]
Figure 2: Image of (Type-α\alpha).
[Uncaptioned image]
Figure 3: Image of (Type-β\beta).

For (Type-α\alpha), we have four cases

  1. (Type-1)

    m1<m2<ml(𝚜)Nsm_{1}<m_{2}<m_{l(\mathtt{s})}\leq Ns

  2. (Type-2)

    m1<m2Ns<ml(𝚜)m_{1}<m_{2}\leq Ns<m_{l(\mathtt{s})}

  3. (Type-3)

    m1Ns<m2<ml(𝚜)m_{1}\leq Ns<m_{2}<m_{l(\mathtt{s})}

  4. (Type-4)

    Ns<m1<m2<mi(𝚜)Ns<m_{1}<m_{2}<m_{i(\mathtt{s})}

and for (Type-β\beta), we have three cases

  1. (Type-6)

    m1<m2Nsm_{1}<m_{2}\leq Ns

  2. (Type-7)

    m1Ns<m2m_{1}\leq Ns<m_{2}

  3. (Type-8)

    Ns<m1<m2Ns<m_{1}<m_{2}.

Also, we will divide the summation by the first site after NsNs.

Definition 4.8.

For each sequence (𝐦,𝐧)(\mathbf{m},\mathbf{n}) in 𝕋N(𝚜,u,v)\mathbb{T}_{N}(\mathtt{s},u,v), there exists an i(𝐦,𝐧){1,,|𝚜|1}i(\mathbf{m},\mathbf{n})\in\{1,\dots,|\mathtt{s}|-1\} such that one of the following holds:

  1. (s\mathrm{(s}-1)\mathrm{1)}

    mi(𝐦,𝐧)Ns<ni(𝐦,𝐧)m_{i(\mathbf{m},\mathbf{n})}\leq Ns<n_{i(\mathbf{m},\mathbf{n})}

  2. (s\mathrm{(s}-2)\mathrm{2)}

    ni(𝐦,𝐧)1Ns<mi(𝐦,𝐧)n_{i(\mathbf{m},\mathbf{n})-1}\leq Ns<m_{i(\mathbf{m},\mathbf{n})}.

Thus, we have to estimate 14 cases. However, the arguments are essentially the same, so we will deal with the following two cases in the next section: (s\mathrm{(s}-1)\mathrm{1)}×\times(Type-1), and (s\mathrm{(s}-2)\mathrm{2)}×\times(Type-6).

5 Bounds of moments

In this section, we will give upper bounds of (s\mathrm{(s}-1)\mathrm{1)}×\times(Type-1) and (s\mathrm{(s}-2)\mathrm{2)}×\times(Type-6).

To give upper bounds of moments, we use some estimates in [12], where they gave upper bounds of third moments in terms of multivariate integrals. Essentially, the method is the same as the one in [12] but our integrands are complicated.

5.1 (s\mathrm{(s}-1)\mathrm{1)}×\times(Type-1)-case

5.1.1 Change of time variables

At first, we will change of time variables (𝐦,𝐧)(\mathbf{m},\mathbf{n}) as follows.

For a while, we fix i(𝐦,𝐧)=ii(\mathbf{m},\mathbf{n})=i.

We change the time sequence as follows:

{uk=nkmk for 1k|𝚜|ki ui=Nsmiu~i=niNsvk=mk+1nk for 1k|𝚜|1 .\displaystyle\begin{cases}u_{k}=n_{k}-m_{k}&\text{ for $1\leq k\leq|\mathtt{s}|$, $k\not=i$ }\\ u_{i}=Ns-m_{i}\\ \widetilde{u}_{i}=n_{i}-Ns\\ v_{k}=m_{k+1}-n_{k}&\text{ for $1\leq k\leq|\mathtt{s}|-1$ }.\end{cases}

Then, we replace the variables of the summation from (𝐦,𝐧)(\mathbf{m},\mathbf{n}) to (𝐮,𝐯,u~i)=(u1,,u|𝚜|,v1,,v|𝚜|1,u~i)(\mathbf{u},\mathbf{v},\widetilde{u}_{i})=(u_{1},\dots,u_{|\mathtt{s}|},v_{1},\dots,v_{|\mathtt{s}|-1},\widetilde{u}_{i}) and enlarge the range of them as follows

D(i,𝚜):={u1,,ui,v1,,vi1{0,,NT}u~i,ui+1,,u|𝚜|,vi,,v|𝚜|1{0,,(ts)N}}.\displaystyle D(i,\mathtt{s}):=\left\{\begin{matrix}u_{1},\dots,u_{i},v_{1},\dots,v_{i-1}\in\{0,\dots,NT\}\\ \widetilde{u}_{i},u_{i+1},\dots,u_{|\mathtt{s}|},v_{i},\dots,v_{|\mathtt{s}|-1}\in\{0,\dots,(t-s)N\}\end{matrix}\right\}.

Since Ikk2I_{k}\leq k-2 and Ikk1I_{k}^{\prime}\leq k-1, we can see that

2mknIknIk2mknk1nk22vk1+vk2\displaystyle 2m_{k}-n_{I_{k}}-n_{I_{k}^{\prime}}\geq 2m_{k}-n_{k-1}-n_{k-2}\geq 2v_{k-1}+v_{k-2} (5.1)
and
Unkmk={Uukfor 1k|𝚜|kiUui+u~ifor k=i.\displaystyle U_{n_{k}-m_{k}}=\begin{cases}U_{u_{k}}\quad&\textrm{for $1\leq k\leq|\mathtt{s}|$, $k\not=i$}\\ U_{u_{i}+\widetilde{u}_{i}}\quad&\textrm{for $k=i$}.\end{cases} (5.2)

Thus, we have

(s\mathrm{(s}-1)\mathrm{1)}×\times(Type-1)
Cq,2Cq,3e2TλN3k=4l=3k1𝚜𝒫~α|𝚜|=kl(𝚜)=li=l(𝚜)k1𝐮,u~,𝐯D(i,𝚜)j=3jl(𝚜)k(Cq,12vj1+vj21)(j=1,,kjieλujNUuj)eλu~i+uiNUu~i+ui,\displaystyle\leq\frac{C_{q,2}C_{q,3}e^{2T\lambda}}{N^{3}}\sum_{k=4}^{\infty}\sum_{l=3}^{k-1}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{\mathcal{P}}_{\alpha}\\ |\mathtt{s}|=k\\ l(\mathtt{s})=l\end{smallmatrix}}\sum_{i=l(\mathtt{s})}^{k-1}\sum_{\mathbf{u},\widetilde{u},\mathbf{v}\in D(i,\mathtt{s})}\prod_{\begin{smallmatrix}j=3\\ j\not=l(\mathtt{s})\end{smallmatrix}}^{k}\left(\frac{C_{q,1}}{2v_{j-1}+v_{j-2}}\wedge 1\right)\left(\prod_{\begin{smallmatrix}j=1,\dots,k\\ j\not=i\end{smallmatrix}}e^{-\lambda\frac{u_{j}}{N}}U_{u_{j}}\right)e^{-\lambda\frac{\widetilde{u}_{i}+u_{i}}{N}}U_{\widetilde{u}_{i}+u_{i}},

where |𝚜|4|\mathtt{s}|\geq 4 since |𝚜|>l(𝚜)>2|\mathtt{s}|>l(\mathtt{s})>2 for (Type-1).

We use the following result.

Lemma 5.1.

[12, Lemma 5.3 and (5.37)] For each λ1\lambda\geq 1 and T1T\geq 1, there exists a constant 𝚌ϑ<\mathtt{c}_{\vartheta}<\infty such that for any N1N\geq 1

u=1NTeλuNUu𝚌ϑ2+logλ.\displaystyle\sum_{u=1}^{NT}e^{-\lambda\frac{u}{N}}U_{u}\leq\frac{\mathtt{c}_{\vartheta}}{2+\log\lambda}.
Corollary 5.2.

For each T1T\geq 1 and t0t\geq 0, there exists a constant cϑ<c_{\vartheta}<\infty such that for any N1N\geq 1

1Nv=0Ntu=1NTeλu+vNUu+v𝚌ϑt2+logλ.\displaystyle\frac{1}{N}\sum_{v=0}^{Nt}\sum_{u=1}^{NT}e^{-\lambda\frac{u+v}{N}}U_{u+v}\leq\frac{\mathtt{c}_{\vartheta}t}{2+\log\lambda}.

Thus,

(s\mathrm{(s}-1)\mathrm{1)}×\times(Type-1)
Cq,2Cq,3e2Tλ(ts)N2k=4(cϑ2+logλ)kl=3k1𝚜𝒫~α|𝚜|=kl(𝚜)=li=l(𝚜)k𝐯D𝐯(i,𝚜)j=3jl(𝚜)k(Cq,12vj1+vj21),\displaystyle\leq\frac{C_{q,2}C_{q,3}e^{2T\lambda}(t-s)}{N^{2}}\sum_{k=4}^{\infty}\left(\frac{c_{\vartheta}}{2+\log\lambda}\right)^{k}\sum_{l=3}^{k-1}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{\mathcal{P}}_{\alpha}\\ |\mathtt{s}|=k\\ l(\mathtt{s})=l\end{smallmatrix}}\sum_{i=l(\mathtt{s})}^{k}\sum_{\mathbf{v}\in D_{\mathbf{v}}(i,\mathtt{s})}\prod_{\begin{smallmatrix}j=3\\ j\not=l(\mathtt{s})\end{smallmatrix}}^{k}\left(\frac{C_{q,1}}{2v_{j-1}+v_{j-2}}\wedge 1\right), (5.3)

where we set

D𝐯(i,𝚜):={v1,,vi1{0,,NT}vi,,v|𝚜|1{0,,(ts)N}}.\displaystyle D_{\mathbf{v}}(i,\mathtt{s}):=\left\{\begin{matrix}v_{1},\dots,v_{i-1}\in\{0,\dots,NT\}\\ v_{i},\dots,v_{|\mathtt{s}|-1}\in\{0,\dots,(t-s)N\}\end{matrix}\right\}. (5.4)

Now, we focus on

𝐯D𝐯(i,𝚜)j=3jl(𝚜)k(Cq,12vj1+vj21)\displaystyle\sum_{\mathbf{v}\in D_{\mathbf{v}}(i,\mathtt{s})}\prod_{\begin{smallmatrix}j=3\\ j\not=l(\mathtt{s})\end{smallmatrix}}^{k}\left(\frac{C_{q,1}}{2v_{j-1}+v_{j-2}}\wedge 1\right)

which is the summation of the product of k3k-3 terms with respect to k1k-1 variables. By the AM-GM inequality a+b2aba+b\geq 2\sqrt{ab} for a0a\geq 0, b0b\geq 0, it is dominated by

1(2N)k3𝐯D𝐯(i,𝚜)j=1jl(𝚜)2k2(Cq,1vj+1Nvj+1N+vjNN)\displaystyle\frac{1}{(2N)^{k-3}}\sum_{\mathbf{v}\in D_{\mathbf{v}}(i,\mathtt{s})}\prod_{\begin{smallmatrix}j=1\\ j\not=l(\mathtt{s})-2\end{smallmatrix}}^{k-2}\left(\frac{C_{q,1}}{\sqrt{\frac{v_{j+1}}{N}}\sqrt{\frac{v_{j+1}}{N}+\frac{v_{j}}{N}}}\wedge N\right)
Cq,1k3N2[0,s]i1×[0,ts]ki𝑑𝐯j=1jl(𝚜)2k21vj+1vj+1+vj.\displaystyle\leq C_{q,1}^{k-3}N^{2}\int_{[0,s]^{i-1}\times[0,t-s]^{k-i}}d\mathbf{v}\prod_{\begin{smallmatrix}j=1\\ j\not=l(\mathtt{s})-2\end{smallmatrix}}^{k-2}\frac{1}{\sqrt{v_{j+1}}\sqrt{v_{j+1}+v_{j}}}. (5.5)

Now, we integrate it in order from v1v_{1} to vk1v_{k-1}.

Since v1v_{1} appear as 1v1+v2\frac{1}{\sqrt{v_{1}+v_{2}}} in integrand of (5.5),

[0,s]i1×[0,ts]ki𝑑𝐯j=1jl(𝚜)2|𝚜|21vj+1vj+1+vj\displaystyle\int_{[0,s]^{i-1}\times[0,t-s]^{k-i}}d\mathbf{v}\prod_{\begin{smallmatrix}j=1\\ j\not=l(\mathtt{s})-2\end{smallmatrix}}^{|\mathtt{s}|-2}\frac{1}{\sqrt{v_{j+1}}\sqrt{v_{j+1}+v_{j}}}
22T[0,T]i2×[0,ts]ki𝑑𝐯(j=2l(𝚜)31vjvj+vj+1)1vl(𝚜)2\displaystyle\leq 2\sqrt{2T}\int_{[0,T]^{i-2}\times[0,t-s]^{k-i}}d\mathbf{v}\left(\prod_{j=2}^{l(\mathtt{s})-3}\frac{1}{\sqrt{v_{j}}\sqrt{v_{j}+v_{j+1}}}\right)\frac{1}{\sqrt{v_{l(\mathtt{s})-2}}}
1vl(𝚜)1+vl(𝚜)(j=l(𝚜)k21vjvj+vj+1)1vk1.\displaystyle\hskip 50.00008pt\frac{1}{\sqrt{v_{l(\mathtt{s})-1}+v_{l(\mathtt{s})}}}\left(\prod_{j^{\prime}=l(\mathtt{s})}^{k-2}\frac{1}{\sqrt{v_{j^{\prime}}}\sqrt{v_{j^{\prime}}+v_{j^{\prime}+1}}}\right)\frac{1}{\sqrt{v_{k-1}}}.

To estimate the integral in the right-and side, we use the results in [12].

We define

ϕt(0)(u)=1,andϕt(k)(u)=0t1s(s+u)ϕt(k1)(s)𝑑s,for k1 and u,t0.\displaystyle\phi_{t}^{(0)}(u)=1,\quad and\quad\phi_{t}^{(k)}(u)=\int_{0}^{t}\frac{1}{\sqrt{s(s+u)}}\phi_{t}^{(k-1)}(s)ds,\quad\text{for $k\geq 1$ and $u,t\geq 0$}.

Then, it is easy to see that

ϕt(0)(u)=ϕ1(0)(u)\displaystyle\phi_{t}^{(0)}(u)=\phi_{1}^{(0)}(u)
ϕt(1)(u)=011s(s+ut)𝑑s=ϕ1(1)(ut)\displaystyle\phi_{t}^{(1)}(u)=\int_{0}^{1}\frac{1}{\sqrt{s(s+\frac{u}{t})}}ds=\phi_{1}^{(1)}\left(\frac{u}{t}\right)
ϕt(k)(u)=011s(s+ut)ϕt(k1)(u)𝑑s=ϕ1(k)(utk)for all k1.\displaystyle\phi_{t}^{(k)}(u)=\int_{0}^{1}\frac{1}{\sqrt{s(s+\frac{u}{t})}}\phi_{t}^{(k-1)}(u)ds=\phi_{1}^{(k)}\left(\frac{u}{t^{k}}\right)\quad\text{for all $k\geq 1$}.
Lemma 5.3.

[12, Lemma 5.4] For all kk\in\mathbb{N},

ϕ1(k)(v)32ki=0k1i!(12loge2v)i32kev,\displaystyle\phi^{(k)}_{1}(v)\leq 32^{k}\sum_{i=0}^{k}\frac{1}{i!}\left(\frac{1}{2}\log\frac{e^{2}}{v}\right)^{i}\leq 32^{k}\frac{e}{\sqrt{v}},

for all v(0,1)v\in(0,1).

In particular,

ϕT(k)32kpki=0k1i!(12plogTke2v)i(32p)kTk2pe1pv12p\displaystyle\phi_{T}^{(k)}\leq 32^{k}p^{k}\sum_{i=0}^{k}\frac{1}{i!}\left(\frac{1}{2p}\log\frac{T^{k}e^{2}}{v}\right)^{i}\leq(32p)^{k}T^{\frac{k}{2p}}e^{\frac{1}{p}}v^{-\frac{1}{2p}}

for T>0T>0 and p1p\geq 1.

Therefore,

[0,T]l(𝚜)3j=2l(𝚜)21vjvj+vj+11vl(𝚜)2dv1dvl(𝚜)2\displaystyle\int_{[0,T]^{l(\mathtt{s})-3}}\prod_{j=2}^{l(\mathtt{s})-2}\frac{1}{\sqrt{v_{j}}\sqrt{v_{j}+v_{j+1}}}\frac{1}{\sqrt{v_{l(\mathtt{s})-2}}}dv_{1}\dots dv_{l(\mathtt{s})-2}
(32p)l(𝚜)3Tl(𝚜)32pe1p0T1v12+12p𝑑v11212p(32p)l(𝚜)3Tl(𝚜)32p+1212pe1p.\displaystyle\leq(32p)^{l(\mathtt{s})-3}T^{\frac{l(\mathtt{s})-3}{2p}}e^{\frac{1}{p}}\int_{0}^{T}\frac{1}{v^{\frac{1}{2}+\frac{1}{2p}}}dv\leq\frac{1}{\frac{1}{2}-\frac{1}{2p}}(32p)^{l(\mathtt{s})-3}T^{\frac{l(\mathtt{s})-3}{2p}+\frac{1}{2}-\frac{1}{2p}}e^{\frac{1}{p}}. (5.6)

for p>1p>1.

Also, we have

[0,T]il(𝚜)+1×[0,ts]ki1vl(𝚜)1+vl(𝚜)(j=l(𝚜)k21vjvj+vj+1)1vk1𝑑vl(𝚜)1𝑑vk1\displaystyle\int_{[0,T]^{i-l(\mathtt{s})+1}\times[0,t-s]^{k-i}}\frac{1}{\sqrt{v_{l(\mathtt{s})-1}+v_{l(\mathtt{s})}}}\left(\prod_{j^{\prime}=l(\mathtt{s})}^{k-2}\frac{1}{\sqrt{v_{j^{\prime}}}\sqrt{v_{j^{\prime}}+v_{j^{\prime}+1}}}\right)\frac{1}{\sqrt{v_{k-1}}}dv_{l(\mathtt{s})-1}\dots dv_{k-1}
2T(32p)kl(𝚜)1Tkl(𝚜)12pe1p0ts1v12+12p𝑑v\displaystyle\leq 2\sqrt{T}(32p)^{k-l(\mathtt{s})-1}T^{\frac{k-l(\mathtt{s})-1}{2p}}e^{\frac{1}{p}}\int_{0}^{t-s}\frac{1}{v^{\frac{1}{2}+\frac{1}{2p}}}dv
21212pT(32p)kl(𝚜)1Tkl(𝚜)12pe1p(ts)1212p.\displaystyle\leq\frac{2}{\frac{1}{2}-\frac{1}{2p}}\sqrt{T}(32p)^{k-l(\mathtt{s})-1}T^{\frac{k-l(\mathtt{s})-1}{2p}}e^{\frac{1}{p}}(t-s)^{\frac{1}{2}-\frac{1}{2p}}. (5.7)

Combining (5.3), (5.5), (5.6), and (5.7), we obtain that

(s-1)×(Type-1)CpCq,2Cq,3e2λT(ts)3212pk=4(cϑ2+logλ)kl=3k1𝚜𝒫~α|𝚜|=kl(𝚜)=li=l(𝚜)k(32p)kl4Tk52p+32\displaystyle\textrm{\ref{item:s-1}$\times$\ref{A-1}}\leq C_{p}C_{q,2}C_{q,3}e^{2\lambda T}(t-s)^{\frac{3}{2}-\frac{1}{2p}}\sum_{k=4}^{\infty}\left(\frac{c_{\vartheta}}{2+\log\lambda}\right)^{k}\sum_{l=3}^{k-1}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{\mathcal{P}}_{\alpha}\\ |\mathtt{s}|=k\\ l(\mathtt{s})=l\end{smallmatrix}}\sum_{i=l(\mathtt{s})}^{k}(32p)^{k-l-4}T^{\frac{k-5}{2p}+\frac{3}{2}}
CpCq,2Cq,3e2λT(ts)3212pk=2(cϑ2+logλ)kk6k(32p)k3Tk52p+32\displaystyle\leq C_{p}C_{q,2}C_{q,3}e^{2\lambda T}(t-s)^{\frac{3}{2}-\frac{1}{2p}}\sum_{k=2}^{\infty}\left(\frac{c_{\vartheta}}{2+\log\lambda}\right)^{k}k6^{k}(32p)^{k-3}T^{\frac{k-5}{2p}+\frac{3}{2}}
Cp,T,λ(ts)3212p\displaystyle\leq C_{p,T,\lambda}(t-s)^{\frac{3}{2}-\frac{1}{2p}}

for λ\lambda large enough, where CpC_{p} is a constant depending only on p>1p>1 and Cp,T,λC_{p,T,\lambda} is a constant depending on p,T,λp,T,\lambda.

5.2 (s\mathrm{(s}-2)\mathrm{2)}×\times(Type-2)-case

For a while, we fix i(𝐦,𝐧)=ii(\mathbf{m},\mathbf{n})=i.

We change the time sequence as follows:

{uk=nkmk, for 1k|𝚜| vk=mk+1nkfor 1k|𝚜|1ki1vi1=Nsni1v~i1=miNs.\displaystyle\begin{cases}u_{k}=n_{k}-m_{k},&\text{ for $1\leq k\leq|\mathtt{s}|$ }\\ v_{k}=m_{k+1}-n_{k}&\text{for $1\leq k\leq|\mathtt{s}|-1$, $k\not=i-1$}\\ v_{i-1}=Ns-n_{i-1}\\ \widetilde{v}_{i-1}=m_{i}-Ns.\end{cases}

Then, we replace the variables of the summation from (𝐦,𝐧)(\mathbf{m},\mathbf{n}) to (𝐮,𝐯,v~i1)(\mathbf{u},\mathbf{v},\widetilde{v}_{i-1}) and enlarge the range of them as follows:

D2(i,𝚜):={u1,,u|𝚜|,v1,,vi1{0,,NT}v~i1,vi,,v|𝚜|1{0,,(ts)N}}.\displaystyle D_{2}(i,\mathtt{s}):=\left\{\begin{matrix}u_{1},\dots,u_{|\mathtt{s}|},v_{1},\dots,v_{i-1}\in\{0,\dots,NT\}\\ \widetilde{v}_{i-1},v_{i},\dots,v_{|\mathtt{s}|-1}\in\{0,\dots,(t-s)N\}\end{matrix}\right\}.

Using (5.1), we can see that

(s-2)×(Type-6)Cq,32e2TλN2k=4𝚜𝒫~β|𝚜|=ki=3k1(𝐮,𝐯,v~i1)D2(i,𝚜)j=3,,k(Cq,12vj1+vIj1)j=1|𝚜|eλujNUuj\displaystyle\textrm{\ref{item:s-2}$\times$\ref{B-1}}\leq\frac{C_{q,3}^{2}e^{2T\lambda}}{N^{2}}\sum_{k=4}^{\infty}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{\mathcal{P}}_{\beta}\\ |\mathtt{s}|=k\end{smallmatrix}}\sum_{i=3}^{k-1}\sum_{\begin{smallmatrix}(\mathbf{u},\mathbf{v},\widetilde{v}_{i-1})\in D_{2}(i,\mathtt{s})\end{smallmatrix}}\prod_{\begin{smallmatrix}j=3,\dots,k\end{smallmatrix}}\left(\frac{C_{q,1}}{2v_{j-1}+v_{I_{j}}}\wedge 1\right)\prod_{j=1}^{|\mathtt{s}|}e^{-\lambda\frac{u_{j}}{N}}U_{u_{j}}
Cq,32e2TλN2k=4𝚜𝒫~β|𝚜|=ki=3k1(𝐮,𝐯,v~i1)D2(i,𝚜)j=3,,k(Cq,12vj1+vj21)j=1|𝚜|eλujNUuj,\displaystyle\leq\frac{C_{q,3}^{2}e^{2T\lambda}}{N^{2}}\sum_{k=4}^{\infty}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{\mathcal{P}}_{\beta}\\ |\mathtt{s}|=k\end{smallmatrix}}\sum_{i=3}^{k-1}\sum_{\begin{smallmatrix}(\mathbf{u},\mathbf{v},\widetilde{v}_{i-1})\in D_{2}(i,\mathtt{s})\end{smallmatrix}}\prod_{\begin{smallmatrix}j=3,\dots,k\end{smallmatrix}}\left(\frac{C_{q,1}}{2v_{j-1}+v_{j-2}}\wedge 1\right)\prod_{j=1}^{|\mathtt{s}|}e^{-\lambda\frac{u_{j}}{N}}U_{u_{j}}, (5.8)

where we remark that i(𝐦,𝐧)|𝚜|1i(\mathbf{m},\mathbf{n})\leq|\mathtt{s}|-1 since m|𝚜|1m_{|\mathtt{s}|-1} must be larger than NsNs. Also, we remark that the summand does not contain v~i1\widetilde{v}_{i-1}.

Lemma 5.1 yields that

(s-2)×(Type-6)Cq,32e2Tλ(ts)Nk=4(cϑ2+logλ)k𝚜𝒫~β|𝚜|=ki=3k(𝐯)D𝐯(i,𝚜)j=3,,k(Cq,12vj+vIj+11)\displaystyle\textrm{\ref{item:s-2}$\times$\ref{B-1}}\leq\frac{C_{q,3}^{2}e^{2T\lambda}(t-s)}{N}\sum_{k=4}^{\infty}\left(\frac{c_{\vartheta}}{2+\log\lambda}\right)^{k}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{\mathcal{P}}_{\beta}\\ |\mathtt{s}|=k\end{smallmatrix}}\sum_{i=3}^{k}\sum_{\begin{smallmatrix}(\mathbf{v})\in D_{\mathbf{v}}(i,\mathtt{s})\end{smallmatrix}}\prod_{\begin{smallmatrix}j=3,\dots,k\end{smallmatrix}}\left(\frac{C_{q,1}}{2v_{j}+v_{I_{j}+1}}\wedge 1\right)

where D𝐯(i,𝚜)D_{\mathbf{v}}(i,\mathtt{s}) is defined in (5.4).

Then, a similar argument to the analysis of (5.5) yields that

𝐯D𝐯(i,𝚜)j=3,,k(Cq,12vj+vIj+11)Cq,1k2N[0,s]i1×[0,ts]ki𝑑𝐯j=1k21vj+1vj+1+vj.\displaystyle\sum_{\begin{smallmatrix}\mathbf{v}\in D_{\mathbf{v}}(i,\mathtt{s})\end{smallmatrix}}\prod_{\begin{smallmatrix}j=3,\dots,k\end{smallmatrix}}\left(\frac{C_{q,1}}{2v_{j}+v_{I_{j}+1}}\wedge 1\right)\leq C_{q,1}^{k-2}N\int_{[0,s]^{i-1}\times[0,t-s]^{k-i}}d\mathbf{v}\prod_{j=1}^{k-2}\frac{1}{\sqrt{v_{j+1}}\sqrt{v_{j+1}+v_{j}}}.

The rest of analysis is almost the same as the one of the proof after (5.5), so we omit it.

Anyway, we can obtain that

(s-2)×(Type-6)Cp,T,λ(ts)3212p\displaystyle\textrm{\ref{item:s-2}$\times$\ref{B-1}}\leq C_{p,T,\lambda}(t-s)^{\frac{3}{2}-\frac{1}{2p}}

for λ\lambda large enough, where Cp,T,λC_{p,T,\lambda} is a constant depending on p,T,λp,T,\lambda.

6 Proof of Lemma 3.14

As we mentioned in Remark 3.18, it is enough to focus on (3.22)-(3.24).

The limit of the third moment of Z¯N;sϕ(ψ)\overline{Z}^{\phi}_{N;s}(\psi) was obtained in [12] and the higher moments of the moments in the continuous setting (stochastic heat equation) was obtained in [25].

We recall that for each 𝚜𝒫~f\mathtt{s}\in\widetilde{\mathcal{P}}_{f} and E{A,B,C,D}E\in\{A,B,C,D\}, we set

i1E=inf{j1:sjE},\displaystyle i^{E}_{1}=\inf\{j\geq 1:s_{j}\ni E\},
and if ikE<,ik+1E=inf{j>ikE:sjE},\displaystyle\text{and if $i^{E}_{k}<\infty$},i_{k+1}^{E}=\inf\{j>i_{k}^{E}:s_{j}\ni E\},

where we set inf=\inf=\infty. Also, we denote by kE=sup{k:ik+1E=}k^{E}=\sup\{k:i_{k+1}^{E}=\infty\} the number of times EE appears in 𝚜\mathtt{s}.

Also, we define

ΘE(𝐮,𝐯,𝐱,𝐲)=Φu1E(x1E)j=1kE1puj+1EvjE(xj+1EyjE)\displaystyle\Theta^{E}(\mathbf{u},\mathbf{v},\mathbf{x},\mathbf{y})=\Phi_{u^{E}_{1}}\left(x^{E}_{1}\right)\prod_{j=1}^{k^{E}-1}p_{u^{E}_{j+1}-v^{E}_{j}}\left(x^{E}_{j+1}-y^{E}_{j}\right)

for ϕCc(2)\phi\in C_{c}({\mathbb{R}}^{2}), 0<u1<v1<<uk<vk0<u_{1}<v_{1}<\dots<u_{k}<v_{k}, x1,y1,,xk,yk2x_{1},y_{1},\dots,x_{k},y_{k}\in{\mathbb{R}}^{2}, and 𝚜𝒫~f\mathtt{s}\in\widetilde{\mathcal{P}}_{f}, where we set ujE=uijEu^{E}_{j}=u_{i^{E}_{j}}, vjE=vijEv^{E}_{j}=v_{i^{E}_{j}}, xjE=xijEx^{E}_{j}=x_{i^{E}_{j}}, and yjE=yijEy^{E}_{j}=y_{i^{E}_{j}} for j=1,,kEj=1,\dots,k^{E} and E{A,B,C,D}E\in\{A,B,C,D\}.

The limits of (3.22)-(3.24) are given as follows.

Lemma 6.1.

Let ϕCc(2)\phi\in C_{c}({\mathbb{R}}^{2}) and ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}). For each t0t\geq 0,

limNE[(0NtN2𝖹¯N;sϕ(pε(z))2ψ(z)2dz)2]\displaystyle\lim_{N\to\infty}E\left[\left(\int_{0}^{\frac{\lfloor Nt\rfloor}{N}}\int_{{\mathbb{R}}^{2}}\overline{\mathsf{Z}}^{\phi}_{N;s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z\right)^{2}\right]
=[0,t]2i=12(Φsi+ε(y)2ψ(y)2dy)ds1ds2\displaystyle=\int_{[0,t]^{2}}\prod_{i=1}^{2}\left(\int\Phi_{s_{i}+\varepsilon}(y)^{2}\psi(y)^{2}\text{\rm d}y\right)\text{\rm d}s_{1}\text{\rm d}s_{2}
+k1𝚜P~f|𝚜|=k(4π)k0<u1<v1<<uk1<vk1<uk<vk<tvkAAvkBB<σ<t,vkCCvkDD<τ<td𝐮d𝐯dσdτ(2)2k+2d𝐱d𝐲dzABdzCD1{zAB=zA=ZB,zCD=zC=zD,σ=σA=σB,τ=σC=σD}\displaystyle+\sum_{k\geq 1}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{P}_{f}\\ |\mathtt{s}|=k\end{smallmatrix}}(4\pi)^{k}\iint_{\begin{smallmatrix}0<u_{1}<v_{1}<\dots<u_{k-1}<v_{k-1}<u_{k}<v_{k}<t\\ v^{A}_{k^{A}}\vee v^{B}_{k^{B}}<\sigma<t,v^{C}_{k^{C}}\vee v^{D}_{k^{D}}<\tau<t\end{smallmatrix}}\text{\rm d}\mathbf{u}\text{\rm d}\mathbf{v}\text{\rm d}\sigma\text{\rm d}\tau\int_{\left({\mathbb{R}}^{2}\right)^{2k+2}}\text{\rm d}\mathbf{x}\text{\rm d}\mathbf{y}\text{\rm d}z^{AB}\text{\rm d}z^{CD}1\left\{\begin{smallmatrix}z^{AB}=z^{A}=Z^{B},z^{CD}=z^{C}=z^{D},\\ \sigma=\sigma^{A}=\sigma^{B},\tau=\sigma^{C}=\sigma^{D}\end{smallmatrix}\right\}
i=1kGϑ(vjuj,yjxj)E{A,B,C,D}ΘE(𝐮,𝐯,𝐱,𝐲)pσEvkEE+ε(zEykEE)ψ(zE)\displaystyle\hskip 30.00005pt\prod_{i=1}^{k}G_{\vartheta}(v_{j}-u_{j},y_{j}-x_{j})\prod_{E\in\{A,B,C,D\}}\Theta^{E}(\mathbf{u},\mathbf{v},\mathbf{x},\mathbf{y})p_{\sigma^{E}-v^{E}_{k^{E}}+\varepsilon}(z^{E}-y^{E}_{k^{E}})\psi(z^{E}) (6.1)
Lemma 6.2.

Let ϕCc(2)\phi\in C_{c}({\mathbb{R}}^{2}) and ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}). For each t0t\geq 0,

limNE[0NtN2𝖹¯N;sϕ(pε(z))2ψ(z)2dzMN,ϕ(ψ)t]\displaystyle\lim_{N\to\infty}E\left[\int_{0}^{\frac{\lfloor Nt\rfloor}{N}}\int_{{\mathbb{R}}^{2}}\overline{\mathsf{Z}}^{\phi}_{N;s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z\left\langle M^{N,\phi}(\psi)\right\rangle_{t}\right]
=k1𝚜P~f|𝚜|=k,sk=CD(4π)k0<u1<v1<<uk1<vk1<uk<vk<tvk1<σ<td𝐮d𝐯dσ(2)2k+1d𝐱d𝐲dzAB1{zAB=zA=ZB}\displaystyle=\sum_{k\geq 1}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{P}_{f}\\ |\mathtt{s}|=k,s_{k}={CD}\end{smallmatrix}}(4\pi)^{k}\iint_{\begin{smallmatrix}0<u_{1}<v_{1}<\dots<u_{k-1}<v_{k-1}<u_{k}<v_{k}<t\\ v_{k-1}<\sigma<t\end{smallmatrix}}\text{\rm d}\mathbf{u}\text{\rm d}\mathbf{v}\text{\rm d}\sigma\int_{\left({\mathbb{R}}^{2}\right)^{2k+1}}\text{\rm d}\mathbf{x}\text{\rm d}\mathbf{y}\text{\rm d}z^{AB}1\left\{z^{AB}=z^{A}=Z^{B}\right\}
i=1kGϑ(vjuj,yjxj)E{A,B,C,D}ΘE(𝐮,𝐯,𝐱,𝐲)F=A,BpσvFkF+ε(zFyFkF)ψ(zF)ψ(yk)2\displaystyle\hskip 30.00005pt\prod_{i=1}^{k}G_{\vartheta}(v_{j}-u_{j},y_{j}-x_{j})\prod_{E\in\{A,B,C,D\}}\Theta^{E}(\mathbf{u},\mathbf{v},\mathbf{x},\mathbf{y})\prod_{F=A,B}p_{\sigma-v^{F}_{k^{F}}+\varepsilon}(z^{F}-y^{F}_{k^{F}})\psi(z^{F})\psi(y_{k})^{2}
+k1𝚜P~f|𝚜|=k,sk1=CD,sk=AB(4π)k0<u1<v1<<uk1<vk1<uk<vk<tvk<σ<td𝐮d𝐯dσ(2)2k+1d𝐱d𝐲\displaystyle+\sum_{k\geq 1}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{P}_{f}\\ |\mathtt{s}|=k,\\ s_{k-1}=CD,s_{k}={AB}\end{smallmatrix}}(4\pi)^{k}\iint_{\begin{smallmatrix}0<u_{1}<v_{1}<\dots<u_{k-1}<v_{k-1}<u_{k}<v_{k}<t\\ v_{k}<\sigma<t\end{smallmatrix}}\text{\rm d}\mathbf{u}\text{\rm d}\mathbf{v}\text{\rm d}\mathbf{\sigma}\int_{\left({\mathbb{R}}^{2}\right)^{2k+1}}\text{\rm d}\mathbf{x}\text{\rm d}\mathbf{y}
i=1kGϑ(viui,yixi)E{A,B,C,D}ΘE(𝐮,𝐯,𝐱,𝐲)pσvk+ε(zyk)2ψ(z)2ψ(yk1)2.\displaystyle\hskip 30.00005pt\prod_{i=1}^{k}G_{\vartheta}(v_{i}-u_{i},y_{i}-x_{i})\prod_{E\in\{A,B,C,D\}}\Theta^{E}(\mathbf{u},\mathbf{v},\mathbf{x},\mathbf{y})p_{\sigma-v_{k}+\varepsilon}\left(z-y_{k}\right)^{2}\psi(z)^{2}\psi(y_{k-1})^{2}. (6.2)
Lemma 6.3.

Let ϕCc(2)\phi\in C_{c}({\mathbb{R}}^{2}) and ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}). For each t0t\geq 0,

limNE[MN,ϕ(ψ)t2]\displaystyle\lim_{N\to\infty}E\left[\langle M^{N,\phi}(\psi)\rangle_{t}^{2}\right]
=2k2𝚜P~f|𝚜|=k,sk1=AB,sk=CD(4π)k0<u1<v1<<uk1<vk1<uk<vk<td𝐮d𝐯(2)2kd𝐱d𝐲\displaystyle=2\sum_{k\geq 2}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{P}_{f}\\ |\mathtt{s}|=k,\\ s_{k-1}=AB,s_{k}=CD\end{smallmatrix}}(4\pi)^{k}\iint_{\begin{smallmatrix}0<u_{1}<v_{1}<\dots<u_{k-1}<v_{k-1}<u_{k}<v_{k}<t\end{smallmatrix}}\text{\rm d}\mathbf{u}\text{\rm d}\mathbf{v}\int_{\left({\mathbb{R}}^{2}\right)^{2k}}\text{\rm d}\mathbf{x}\text{\rm d}\mathbf{y}
i=1kGϑ(viui,yixi)E{A,B,C,D}ΘE(𝐮,𝐯,𝐱,𝐲)ψ(yk1)2ψ(yk)2.\displaystyle\hskip 30.00005pt\prod_{i=1}^{k}G_{\vartheta}(v_{i}-u_{i},y_{i}-x_{i})\prod_{E\in\{A,B,C,D\}}\Theta^{E}(\mathbf{u},\mathbf{v},\mathbf{x},\mathbf{y})\psi(y_{k-1})^{2}\psi(y_{k})^{2}. (6.3)

We give an outline of the proof of Lemma 6.2 and omit the proofs of Lemma 6.1 and 6.3 since the argument are almost the same.

Remark 6.4.

The summations in (6.1)-(6.3) converge absolutely. It follows from the following alternative representation of (2.14) for h=4h=4:

(2.14)\displaystyle\eqref{eq:highermomentsrepre}
=(2)4E{A,B,C,D}ϕ(xE)pt(xE,yE)ψ(yE)d𝐱d𝐲\displaystyle=\int_{({\mathbb{R}}^{2})^{4}}\prod_{E\in\{A,B,C,D\}}\phi(x_{E})p_{t}(x_{E},y_{E})\psi(y_{E})\text{\rm d}\mathbf{x}\text{\rm d}\mathbf{y}
+k1𝚜𝒫~f|𝚜|=k(4π)k∫⋯∫0<u1<v1<<uk<vk<td𝐮d𝐯(𝐑2)2kd𝐱d𝐲\displaystyle+\sum_{k\geq 1}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{\mathcal{P}}_{f}\\ |\mathtt{s}|=k\end{smallmatrix}}(4\pi)^{k}\idotsint\limits_{0<u_{1}<v_{1}<\dots<u_{k}<v_{k}<t}\text{\rm d}\mathbf{u}\text{\rm d}\mathbf{v}\int_{(\mathbf{R}^{2})^{2k}}\text{\rm d}\mathbf{x}\text{\rm d}\mathbf{y}
i=1kGϑ(viui,yixi)E{A,B,C,D}ΘE(𝐮,𝐯,𝐱,𝐲)E{A,B,C,D}ΨtvEkE(yEkE).\displaystyle\hskip 40.00006pt\prod_{i=1}^{k}G_{\vartheta}(v_{i}-u_{i},y_{i}-x_{i})\prod_{E\in\{A,B,C,D\}}\Theta^{E}(\mathbf{u},\mathbf{v},\mathbf{x},\mathbf{y})\prod_{E\in\{A,B,C,D\}}\Psi_{t-v^{E}_{k_{E}}}(y^{E}_{k^{E}}). (6.4)

In (6.1), we may take ψ1\psi\equiv 1 and ϕ0\phi\geq 0. Also, we have vk1<σ<tdσ2pσvEkA+ε(zyAkA)pσvEkB+ε(zyBkB)dzCε\int_{v_{k-1}<\sigma<t}\text{\rm d}\sigma\int_{{\mathbb{R}}^{2}}p_{\sigma-v^{E}_{k^{A}}+\varepsilon}(z-y^{A}_{k^{A}})p_{\sigma-v^{E}_{k^{B}}+\varepsilon}(z-y^{B}_{k^{B}})\text{\rm d}z\leq C_{\varepsilon} uniformly in vk1v_{k-1}, ε>0\varepsilon>0, and x,y2x,y\in{\mathbb{R}}^{2}. Then, this upper bound has the same form as (6.4).

6.1 Proof of Lemma 6.2

We first give the chaos expansion of the expectation of

E[(0NtN2𝖹¯ϕN;s(pε(z))2ψ(z)2dz)MN,ϕ(ψ)t]\displaystyle E\left[\left(\int_{0}^{\frac{\lfloor Nt\rfloor}{N}}\int_{{\mathbb{R}}^{2}}\overline{\mathsf{Z}}^{\phi}_{N;s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z\right)\left\langle M^{N,\phi}(\psi)\right\rangle_{t}\right]

in a similar way to the argument in Section 4. We use the label A,BA,B derived from the “random walks” in 0NtN2𝖹¯ϕN;s(pε(z))2ψ(z)2dz\int_{0}^{\frac{\lfloor Nt\rfloor}{N}}\int_{{\mathbb{R}}^{2}}\overline{\mathsf{Z}}^{\phi}_{N;s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z and C,DC,D derived from the “random walks” in MN,ϕ(ψ)t\left\langle M^{N,\phi}(\psi)\right\rangle_{t}.

We remark that after the last intersection between AA and BB, each of them may meet CC or DD but after the last intersection between CC and DD, neither CC nor DD will meet other particles. Thus, 𝚜𝒫f\mathtt{s}\in\mathcal{P}_{f} contributing the chaos expansion should satisfy one of

  1. (1)

    𝚜|𝚜|=CD\mathtt{s}_{|\mathtt{s}|}=CD

  2. (2)

    𝚜|𝚜|1=CD\mathtt{s}_{|\mathtt{s}|-1}=CD and 𝚜|𝚜|=AB\mathtt{s}_{|\mathtt{s}|}=AB.

As we mentioned in Remark 4.4, quadruple intersections in the chaos expansion of moments are negligible. Therefore, we have the following representation of the moment. We omit its proof since it is almost the same as the discussion in (4.9).

Lemma 6.5.

Let ϕCc(2)\phi\in C_{c}({\mathbb{R}}^{2}) and ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}). For each t0t\geq 0, we have

E[(0NtN2𝖹¯ϕN;s(pε(z))2ψ(z)2dz)MN,ϕ(ψ)t]\displaystyle E\left[\left(\int_{0}^{\frac{\lfloor Nt\rfloor}{N}}\int_{{\mathbb{R}}^{2}}\overline{\mathsf{Z}}^{\phi}_{N;s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z\right)\left\langle M^{N,\phi}(\psi)\right\rangle_{t}\right]
=1N5k1𝚜𝒫f|𝚜|=k,sk=CDx1,y1,xk,yk(m1,n1,,mk,nk)TˇN(𝚜,u,v)nk1uNt\displaystyle=\frac{1}{N^{5}}\sum_{k\geq 1}\sum_{\begin{smallmatrix}\mathtt{s}\in\mathcal{P}_{f}\\ |\mathtt{s}|=k,s_{k}=CD\end{smallmatrix}}\sum_{\begin{smallmatrix}x_{1},y_{1}\dots,x_{k},y_{k}\end{smallmatrix}}\sum_{\begin{smallmatrix}(m_{1},n_{1},\dots,m_{k},n_{k})\in\widecheck{T}_{N}(\mathtt{s},u,v)\\ n_{k-1}\leq u\leq Nt\end{smallmatrix}}
i=1nU(nimi,yixi)E{A,B,C,D}Θ~(N)(𝐦,𝐧,𝐱,𝐲)ψN(yk)2dzQA(𝐲,u,ε,ψ,z)QB(𝐲,u,ε,ψ,z)\displaystyle\hskip 40.00006pt\prod_{i=1}^{n}U(n_{i}-m_{i},y_{i}-x_{i})\prod_{E\in\{A,B,C,D\}}\widetilde{\Theta}^{(N)}(\mathbf{m},\mathbf{n},\mathbf{x},\mathbf{y})\psi_{N}(y_{k})^{2}\int\text{\rm d}zQ_{A}(\mathbf{y},u,\varepsilon,\psi,z)Q_{B}(\mathbf{y},u,\varepsilon,\psi,z)
+1N5k1𝚜𝒫f|𝚜|=k,sk1=CD,sk=ABx1,y1,,xk,yk(m1,n1,,mk,nk)TˇN(𝚜,u,v)nk<uNt\displaystyle+\frac{1}{N^{5}}\sum_{k\geq 1}\sum_{\begin{smallmatrix}\mathtt{s}\in\mathcal{P}_{f}\\ |\mathtt{s}|=k,s_{k-1}=CD,s_{k}=AB\end{smallmatrix}}\sum_{x_{1},y_{1},\dots,x_{k},y_{k}}\sum_{\begin{smallmatrix}(m_{1},n_{1},\dots,m_{k},n_{k})\in\widecheck{T}_{N}(\mathtt{s},u,v)\\ n_{k}<u\leq Nt\end{smallmatrix}}
i=1nU(nimi,yixi)E{A,B,C,D}Θ~(N)(𝐦,𝐧,𝐱,𝐲)ψN(yk)2QA(𝐲,u,ε,ψ,z)QB(𝐲,u,ε,ψ,z)+o(1)\displaystyle\hskip 40.00006pt\prod_{i=1}^{n}U(n_{i}-m_{i},y_{i}-x_{i})\prod_{E\in\{A,B,C,D\}}\widetilde{\Theta}^{(N)}(\mathbf{m},\mathbf{n},\mathbf{x},\mathbf{y})\psi_{N}(y_{k})^{2}Q_{A}(\mathbf{y},u,\varepsilon,\psi,z)Q_{B}(\mathbf{y},u,\varepsilon,\psi,z)+o(1) (6.5)

where TˇN(u,v)\widecheck{T}_{N}(u,v) is the set of time sequences (m1,n1,,m|𝚜|,n|𝚜|)(m_{1},n_{1},\dots,m_{|\mathtt{s}|},n_{|\mathtt{s}|}) which satisfy the followings:

  1. (Tˇ\widecheck{T}-1)

    mi,nim_{i},n_{i} are associated with the stretch sis_{i} for 1i|𝚜|1\leq i\leq|\mathtt{s}|, which represents the start time and the end time of the stretch.

  2. (Tˇ\widecheck{T}-2)

    1m1n1m2n2m|𝚜|n|𝚜|1\leq m_{1}\leq n_{1}\leq m_{2}\leq n_{2}\leq\dots\leq m_{|\mathtt{s}|}\leq n_{|\mathtt{s}|}.

  3. (Tˇ\widecheck{T}-3)

    If sis_{i} and si+1s_{i+1} are not a couple, then ni<mi+1n_{i}<m_{i+1}. Otherwise, ni=mi+1n_{i}=m_{i+1} is allowed.

Also, we set

Θ~(N)(𝐦,𝐧,𝐱,𝐲)=qmE1(ϕN,x1E)j=1kE1qmj+1EnjE(yEj,xEj+1)\displaystyle\widetilde{\Theta}^{(N)}(\mathbf{m},\mathbf{n},\mathbf{x},\mathbf{y})=q_{m^{E}_{1}}(\phi_{N},x_{1}^{E})\prod_{j=1}^{k^{E}-1}q_{m_{j+1}^{E}-n_{j}^{E}}\left(y^{E}_{j},x^{E}_{j+1}\right)
QE(𝐲,u,ε,ψ,z)=y2qunEkE(yEkE,y)pε(yNz)ψ(z)for E=A,B,\displaystyle Q_{E}(\mathbf{y},u,\varepsilon,\psi,z)=\sum_{y\in{\mathbb{Z}}^{2}}q_{u-n^{E}_{k^{E}}}(y^{E}_{k^{E}},y)p_{\varepsilon}\left(\frac{y}{\sqrt{N}}-z\right)\psi(z)\quad\text{for $E=A,B$},

where we write mEj=miEjm^{E}_{j}=m_{i^{E}_{j}}, nEj=niEjn^{E}_{j}=n_{i^{E}_{j}}, xEj=xiEjx^{E}_{j}=x_{i^{E}_{j}}, and yEj=yiEjy^{E}_{j}=y_{i^{E}_{j}}.

Thus, it is enough to see the limit of the right-hand side of (6.5) for the proof of Lemma 6.2. Here, we give an idea of the proof of this convergence since it is almost the same as the proof of [12, (5.3)].

Indeed, we may regard it as “Riemannian summation” for some function due to the following approximation:

qn(ϕN,x)2ϕ(y)pnN(xNy)dy\displaystyle q_{n}(\phi_{N},x)\sim\int_{{\mathbb{R}}^{2}}\phi(y)p_{\frac{n}{N}}\left(\frac{x}{\sqrt{N}}-y\right)\text{\rm d}y
y2qn(x,y)pε(yNz)ψ(z)pnN+ε(zxN)ψ(z)\displaystyle\sum_{y\in{\mathbb{Z}}^{2}}q_{n}(x,y)p_{\varepsilon}\left(\frac{y}{\sqrt{N}}-z\right)\psi(z)\sim p_{\frac{n}{N}+\varepsilon}\left(z-\frac{x}{\sqrt{N}}\right)\psi(z)
qn(x,y)1NpnN(yxN)\displaystyle q_{n}(x,y)\sim\frac{1}{N}p_{\frac{n}{N}}\left(\frac{y-x}{\sqrt{N}}\right)
UN(n,z)4πN2Gϑ(nN,xN).\displaystyle U_{N}\left(n,z\right)\sim\frac{4\pi}{N^{2}}G_{\vartheta}\left(\frac{n}{N},\frac{x}{\sqrt{N}}\right).

In particular, we can find that the approximations of EΘ~(N)(𝐦,𝐧,𝐱,𝐲)\prod_{E}\widetilde{\Theta}^{(N)}(\mathbf{m},\mathbf{n},\mathbf{x},\mathbf{y}) and i=1nU(nimi,yixi)\prod_{i=1}^{n}U(n_{i}-m_{i},y_{i}-x_{i}) yield the factor 1NE{A,B,C,D}(kE1)=1N2k4\frac{1}{N^{\sum_{E\in\{A,B,C,D\}}(k^{E}-1)}}=\frac{1}{N^{2k-4}} and 1N2k\frac{1}{N^{2k}}, respectively, so we obtain the factor 1N4k+1\frac{1}{N^{4k+1}}.

To prove the convergence, we first look at the second term of (6.5) with the near diagonal sets are cut off, i.e.

{nimiεN,(1i|𝚜|),mi+1niεN,(0i|𝚜|1),un|𝚜|>εN}\displaystyle\left\{\begin{array}[]{l}\displaystyle n_{i}-m_{i}\geq\varepsilon N,\quad(1\leq i\leq|\mathtt{s}|),\\ \displaystyle m_{i+1}-n_{i}\geq\varepsilon N,\quad(0\leq i\leq|\mathtt{s}|-1),\\ u-n_{|\mathtt{s}|}>\varepsilon N\end{array}\right\}

for some fixed ε>0\varepsilon>0. It approximates (6.5) uniformly in NN since we know that the boundedness of the fourth moment of 𝖹¯ϕN;s(1)\overline{\mathsf{Z}}^{\phi}_{N;s}(1) and the second moment of MN,ϕ(ψ)t\langle M^{N,\phi}(\psi)\rangle_{t}.

Also, we can find from Remark 6.4 that the second term of (6.2) is approximated by the one restricted by

{viui>ε,(1i|𝚜|),ui+1vi>ε,(0i|𝚜|1),uv|𝚜|>ε}.\displaystyle\left\{\begin{array}[]{l}v_{i}-u_{i}>\varepsilon,\quad(1\leq i\leq|\mathtt{s}|),\\ u_{i+1}-v_{i}>\varepsilon,\quad(0\leq i\leq|\mathtt{s}|-1),\\ u-v_{|\mathtt{s}|}>\varepsilon\end{array}\right\}.

Similarly to the arguments in [12, (5.3)], we need to consider the sumations or integrals with the restricted spatial variables 𝐱\mathbf{x} and 𝐲\mathbf{y} to the set

{|x1|MN,|yixi|MN,(1i|𝚜|),|xi+1yi|MN,(0i|𝚜|1)}\displaystyle\left\{\begin{array}[]{l}|x_{1}|\leq M\sqrt{N},|y_{i}-x_{i}|\leq M\sqrt{N},\quad(1\leq i\leq|\mathtt{s}|),\\ |x_{i+1}-y_{i}|\leq M\sqrt{N},\quad(0\leq i\leq|\mathtt{s}|-1)\end{array}\right\}\quad for (6.2)
{|x1|M,|yixi|M,(1i|𝚜|),|xi+1yi|M,(0i|𝚜|1)}\displaystyle\left\{\begin{array}[]{l}|x_{1}|\leq M,|y_{i}-x_{i}|\leq M,\quad(1\leq i\leq|\mathtt{s}|),\\ |x_{i+1}-y_{i}|\leq M,\quad(0\leq i\leq|\mathtt{s}|-1)\end{array}\right\}\quad for (6.5)

for large M>0M>0.

Proof of Lemma 3.14.

Lemma 6.1-6.3 yield that

lim¯NE[(0NtN(4πlogε2𝖹¯ϕN;s(pε(z))2ψ(z)2dzσN2Ny2Z¯N;Nsϕ(y)2ψN(y)2)ds)2]\displaystyle\varliminf_{N\to\infty}E\left[\left(\int_{0}^{\frac{\lfloor Nt\rfloor}{N}}\left(\frac{4\pi}{-\log\varepsilon}\int_{{\mathbb{R}}^{2}}\overline{\mathsf{Z}}^{\phi}_{N;s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z-{\frac{\sigma_{N}^{2}}{N}\sum_{y\in{\mathbb{Z}}^{2}}{\overline{Z}}_{N;\left\lfloor{Ns}\right\rfloor}^{\phi}(y)^{2}\psi_{N}(y)^{2}}\right)\text{\rm d}s\right)^{2}\right] (6.6)
=(4π)2(logε)2(6.1)24πlogε(6.2)+(6.3).\displaystyle=\frac{(4\pi)^{2}}{(-\log\varepsilon)^{2}}\eqref{eq:momentTT1}-2\frac{4\pi}{-\log\varepsilon}\eqref{eq:momentT1-1}+\eqref{eq:moments111}. (6.7)

Thus, it is enough to see

(4π2)2(logε)2(6.1)24π2logε(6.2)+(6.3)0\displaystyle\frac{(4\pi^{2})^{2}}{(-\log\varepsilon)^{2}}\eqref{eq:momentTT1}-2\frac{4\pi^{2}}{-\log\varepsilon}\eqref{eq:momentT1-1}+\eqref{eq:moments111}\to 0

as ε0\varepsilon\to 0.

For (6.1), we first remark that

1logε|vAkAvBkB<σ<tdσ2dzpσvAkA+ε(zyAkA)pσvBkB+ε(zyBkB)ψ(z)2|\displaystyle\frac{1}{-\log\varepsilon}\left|\int_{v^{A}_{k^{A}}\vee v^{B}_{k^{B}}<\sigma<t}\text{\rm d}\sigma\int_{{\mathbb{R}}^{2}}\text{\rm d}zp_{\sigma-v^{A}_{k^{A}}+\varepsilon}\left(z-y^{A}_{k^{A}}\right)p_{\sigma-v^{B}_{k^{B}}+\varepsilon}\left(z-y^{B}_{k^{B}}\right)\psi(z)^{2}\right|
1logεψvAkAvBkB<σ<tdσp2σvAkAvBkB+2ε(0)ψ22πCt\displaystyle\leq\frac{1}{-\log\varepsilon}\|\psi\|_{\infty}\int_{v^{A}_{k^{A}}\vee v^{B}_{k^{B}}<\sigma<t}\text{\rm d}\sigma p_{2\sigma-v^{A}_{k^{A}}-v^{B}_{k^{B}}+2\varepsilon}(0)\leq\frac{\|\psi\|^{2}_{\infty}}{2\pi}C_{t} (6.8)

for 0<vAkAvBkB<t0<v^{A}_{k^{A}}\vee v^{B}_{k^{B}}<t and some Ct>0C_{t}>0. Moreover, we find that

1logεvAkAvBkB<σ<tdσ2dzpσvAkA+ε(zyAkA)pσvBkB+ε(zyBkB)ψ(z)2\displaystyle\frac{1}{-\log\varepsilon}\int_{v^{A}_{k^{A}}\vee v^{B}_{k^{B}}<\sigma<t}\text{\rm d}\sigma\int_{{\mathbb{R}}^{2}}\text{\rm d}zp_{\sigma-v^{A}_{k^{A}}+\varepsilon}\left(z-y^{A}_{k^{A}}\right)p_{\sigma-v^{B}_{k^{B}}+\varepsilon}\left(z-y^{B}_{k^{B}}\right)\psi(z)^{2}
{14πψ(vAkA)2if vAkA=vBkB and yAkA=yBkBkA=kB0otherwise.\displaystyle\to\begin{cases}\frac{1}{4\pi}\psi(v^{A}_{k^{A}})^{2}\quad&\text{if }v^{A}_{k^{A}}=v^{B}_{k^{B}}\text{ and }y^{A}_{k^{A}}=y^{B}_{k^{B}}\Leftrightarrow\text{$k^{A}=k^{B}$}\\ 0&\text{otherwise}.\end{cases}

by Lemma 3.15. It does hold for C,DC,D.

In particular, they converge to 14πψ(zA)2\frac{1}{4\pi}\psi(z_{A})^{2} and 14πψ(zC)2\frac{1}{4\pi}\psi(z_{C})^{2} if and only if kA=kBk_{A}=k_{B} and kC=kDk_{C}=k_{D} (\Leftrightarrow (kA,kB,kC,kD)=(k1,k1,k,k)(k^{A},k^{B},k^{C},k^{D})=(k-1,k-1,k,k) or (k,k,k1,k1)(k,k,k-1,k-1)).

The dominated convergence theorem yields

1(logε)2(6.1)\displaystyle\frac{1}{(-\log\varepsilon)^{2}}\eqref{eq:momentTT1}
k2𝚜𝒫~f|𝚜|=ksk1=AB,sk=CD(4π)k2∫⋯∫0<u1<v1<<uk<vk<td𝐮d𝐯(𝐑2)2kd𝐱d𝐲\displaystyle\to\sum_{k\geq 2}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{\mathcal{P}}_{f}\\ |\mathtt{s}|=k\\ s_{k-1}=AB,s_{k}=CD\end{smallmatrix}}(4\pi)^{k-2}\idotsint\limits_{0<u_{1}<v_{1}<\dots<u_{k}<v_{k}<t}\text{\rm d}\mathbf{u}\text{\rm d}\mathbf{v}\int_{(\mathbf{R}^{2})^{2k}}\text{\rm d}\mathbf{x}\text{\rm d}\mathbf{y}
i=1kGϑ(viui,yixi)E{A,B,C,D}ΘE(𝐮,𝐯,𝐱,𝐲)E{A,B,C,D}ψ(yk1)2ψ(yk)2\displaystyle\hskip 40.00006pt\prod_{i=1}^{k}G_{\vartheta}(v_{i}-u_{i},y_{i}-x_{i})\prod_{E\in\{A,B,C,D\}}\Theta^{E}(\mathbf{u},\mathbf{v},\mathbf{x},\mathbf{y})\prod_{E\in\{A,B,C,D\}}\psi(y_{k-1})^{2}\psi(y_{k})^{2}
+k2𝚜𝒫~f|𝚜|=ksk1=CD,sk=AB(4π)k2∫⋯∫0<u1<v1<<uk<vk<td𝐮d𝐯(𝐑2)2kd𝐱d𝐲\displaystyle+\sum_{k\geq 2}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{\mathcal{P}}_{f}\\ |\mathtt{s}|=k\\ s_{k-1}=CD,s_{k}=AB\end{smallmatrix}}(4\pi)^{k-2}\idotsint\limits_{0<u_{1}<v_{1}<\dots<u_{k}<v_{k}<t}\text{\rm d}\mathbf{u}\text{\rm d}\mathbf{v}\int_{(\mathbf{R}^{2})^{2k}}\text{\rm d}\mathbf{x}\text{\rm d}\mathbf{y}
i=1kGϑ(viui,yixi)E{A,B,C,D}ΘE(𝐮,𝐯,𝐱,𝐲)E{A,B,C,D}ψ(yk1)2ψ(yk)2\displaystyle\hskip 40.00006pt\prod_{i=1}^{k}G_{\vartheta}(v_{i}-u_{i},y_{i}-x_{i})\prod_{E\in\{A,B,C,D\}}\Theta^{E}(\mathbf{u},\mathbf{v},\mathbf{x},\mathbf{y})\prod_{E\in\{A,B,C,D\}}\psi(y_{k-1})^{2}\psi(y_{k})^{2}
=1(4π)2(6.3)\displaystyle=\frac{1}{(4\pi)^{2}}\eqref{eq:moments111}

Applying the same argument to (6.2), we obtain that

1(logε)(6.2)\displaystyle\frac{1}{(-\log\varepsilon)}\eqref{eq:momentT1-1}
k2𝚜𝒫~f|𝚜|=ksk1=AB,sk=CD(4π)k1∫⋯∫0<u1<v1<<uk<vk<td𝐮d𝐯(𝐑2)2kd𝐱d𝐲\displaystyle\to\sum_{k\geq 2}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{\mathcal{P}}_{f}\\ |\mathtt{s}|=k\\ s_{k-1}=AB,s_{k}=CD\end{smallmatrix}}(4\pi)^{k-1}\idotsint\limits_{0<u_{1}<v_{1}<\dots<u_{k}<v_{k}<t}\text{\rm d}\mathbf{u}\text{\rm d}\mathbf{v}\int_{(\mathbf{R}^{2})^{2k}}\text{\rm d}\mathbf{x}\text{\rm d}\mathbf{y}
i=1kGϑ(viui,yixi)E{A,B,C,D}ΘE(𝐮,𝐯,𝐱,𝐲)E{A,B,C,D}ψ(yk1)2ψ(yk)2\displaystyle\hskip 40.00006pt\prod_{i=1}^{k}G_{\vartheta}(v_{i}-u_{i},y_{i}-x_{i})\prod_{E\in\{A,B,C,D\}}\Theta^{E}(\mathbf{u},\mathbf{v},\mathbf{x},\mathbf{y})\prod_{E\in\{A,B,C,D\}}\psi(y_{k-1})^{2}\psi(y_{k})^{2}
+k2𝚜𝒫~f|𝚜|=ksk1=CD,sk=AB(4π)k1∫⋯∫0<u1<v1<<uk<vk<td𝐮d𝐯(𝐑2)2kd𝐱d𝐲\displaystyle+\sum_{k\geq 2}\sum_{\begin{smallmatrix}\mathtt{s}\in\widetilde{\mathcal{P}}_{f}\\ |\mathtt{s}|=k\\ s_{k-1}=CD,s_{k}=AB\end{smallmatrix}}(4\pi)^{k-1}\idotsint\limits_{0<u_{1}<v_{1}<\dots<u_{k}<v_{k}<t}\text{\rm d}\mathbf{u}\text{\rm d}\mathbf{v}\int_{(\mathbf{R}^{2})^{2k}}\text{\rm d}\mathbf{x}\text{\rm d}\mathbf{y}
i=1kGϑ(viui,yixi)E{A,B,C,D}ΘE(𝐮,𝐯,𝐱,𝐲)E{A,B,C,D}ψ(yk1)2ψ(yk)2\displaystyle\hskip 40.00006pt\prod_{i=1}^{k}G_{\vartheta}(v_{i}-u_{i},y_{i}-x_{i})\prod_{E\in\{A,B,C,D\}}\Theta^{E}(\mathbf{u},\mathbf{v},\mathbf{x},\mathbf{y})\prod_{E\in\{A,B,C,D\}}\psi(y_{k-1})^{2}\psi(y_{k})^{2}
=14π(6.3).\displaystyle=\frac{1}{4\pi}\eqref{eq:moments111}.

Proof of (1.16) in Theorem 1.18.

It is enough to show that

limε0E[(4πlogε0t2𝒵ϑ,ϕs(pε(z))2ψ(z)2dzdsϑ,ϕ(ψ)t)2]=0.\displaystyle\lim_{\varepsilon\to 0}E\left[\left(\frac{4\pi}{-\log\varepsilon}\int_{0}^{t}\int_{{\mathbb{R}}^{2}}\mathscr{Z}^{\vartheta,\phi}_{s}(p_{\varepsilon}(\cdot-z))^{2}\psi(z)^{2}\text{\rm d}z\text{\rm d}s-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}\right)^{2}\right]=0.

Then, we can see that the expectations is give by (6.3) with ψb(2)\psi\in\mathcal{B}_{b}({\mathbb{R}}^{2}). ∎

Remark 6.6.

In Theorem 1.11, quadratic variation is approximated by using 𝒵ϑ,ϕ(pε(x))\mathscr{Z}^{\vartheta,\phi}_{\cdot}(p_{\varepsilon}(\cdot-x)). However, we can approximate it by using 𝒵ϑ,ϕ(1εf(xε))\mathscr{Z}^{\vartheta,\phi}_{\cdot}\left(\frac{1}{\varepsilon}f\left(\frac{\cdot-x}{\sqrt{\varepsilon}}\right)\right) with fCc+(2)f\in C_{c}^{+}({\mathbb{R}}^{2}) satisfying 2f(x)dx=1\int_{{\mathbb{R}}^{2}}f(x)\text{\rm d}x=1.

Indeed, we can modify the proof by using the following lemma instead of Lemma 3.15

Lemma 6.7.

Let ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}), fCc+(2)f\in C_{c}^{+}({\mathbb{R}}^{2}), and T>0T>0. Then, for each 0<t<T0<t<T, there exists CT,f,ψC_{T,f,\psi} such that

sup0<ε<12|1logε2dz0tds2×2dwdwps(wx)ps(wx)1εf(wzε)1εf(wzε)ψ(z)dz|CT,f,ϕ\displaystyle\sup_{0<\varepsilon<\frac{1}{2}}\left|\frac{-1}{\log\varepsilon}\int_{{\mathbb{R}}^{2}}\text{\rm d}z\int_{0}^{t}\text{\rm d}s\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}w\text{\rm d}w^{\prime}p_{s}(w-x)p_{s}(w^{\prime}-x)\frac{1}{\varepsilon}f\left(\frac{w-z}{\sqrt{\varepsilon}}\right)\frac{1}{\varepsilon}f\left(\frac{w^{\prime}-z}{\sqrt{\varepsilon}}\right)\psi(z)\text{\rm d}z\right|\leq C_{T,f,\phi} (6.9)

and

limε01logε2dz0tds2×2dwdwps(wx)ps(wx)1εf(wzε)1εf(wzε)ψ(z)dz=ψ(x)\displaystyle\lim_{\varepsilon\to 0}\frac{-1}{\log\varepsilon}\int_{{\mathbb{R}}^{2}}\text{\rm d}z\int_{0}^{t}\text{\rm d}s\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}w\text{\rm d}w^{\prime}p_{s}(w-x)p_{s}(w^{\prime}-x)\frac{1}{\varepsilon}f\left(\frac{w-z}{\sqrt{\varepsilon}}\right)\frac{1}{\varepsilon}f\left(\frac{w^{\prime}-z}{\sqrt{\varepsilon}}\right)\psi(z)\text{\rm d}z=\psi(x) (6.10)

for each x2x\in{\mathbb{R}}^{2}.

Proof.

(6.9) follows from (3.16) and there exists a constant C>0C>0 such that f(x)Cp1(x)f(x)\leq Cp_{1}(x) for any x2x\in{\mathbb{R}}^{2}.

For (6.10), we will see that

2dz0tds2×2dwdwps(wx)ps(wx)1εf(wzε)1εf(wzε)ψ(z)dz\displaystyle\int_{{\mathbb{R}}^{2}}\text{\rm d}z\int_{0}^{t}\text{\rm d}s\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}w\text{\rm d}w^{\prime}p_{s}(w-x)p_{s}(w^{\prime}-x)\frac{1}{\varepsilon}f\left(\frac{w-z}{\sqrt{\varepsilon}}\right)\frac{1}{\varepsilon}f\left(\frac{w^{\prime}-z}{\sqrt{\varepsilon}}\right)\psi(z)\text{\rm d}z
=2dz0tds2×2dydyps(z+εyx)ps(z+εyx)f(y)f(y)ψ(z)dz\displaystyle=\int_{{\mathbb{R}}^{2}}\text{\rm d}z\int_{0}^{t}\text{\rm d}s\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}y\text{\rm d}y^{\prime}p_{s}(z+\sqrt{\varepsilon}y-x)p_{s}(z+\sqrt{\varepsilon}y^{\prime}-x)f\left(y\right)f\left(y^{\prime}\right)\psi(z)\text{\rm d}z
=ε2dz~0tds2×2dydyps(ε(y+z~))ps(ε(y+z~))f(y)f(y)ψ(x+εz~)dz~\displaystyle=\varepsilon\int_{{\mathbb{R}}^{2}}\text{\rm d}\widetilde{z}\int_{0}^{t}\text{\rm d}s\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}y\text{\rm d}y^{\prime}p_{s}(\sqrt{\varepsilon}(y+\widetilde{z}))p_{s}(\sqrt{\varepsilon}(y^{\prime}+\widetilde{z}))f\left(y\right)f\left(y^{\prime}\right)\psi(x+\sqrt{\varepsilon}\widetilde{z})\text{\rm d}\widetilde{z}
=2dz~0tεdu2×2dydypu(y+z~)pu(y+z~)f(y)f(y)ψ(x+εz~)dz~\displaystyle=\int_{{\mathbb{R}}^{2}}\text{\rm d}\widetilde{z}\int_{0}^{\frac{t}{\varepsilon}}\text{\rm d}u\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}y\text{\rm d}y^{\prime}p_{u}(y+\widetilde{z})p_{u}(y^{\prime}+\widetilde{z})f\left(y\right)f\left(y^{\prime}\right)\psi(x+\sqrt{\varepsilon}\widetilde{z})\text{\rm d}\widetilde{z}
=2dz~0tεdu2×2dydypu(y+z~)pu(y+z~)f(y)f(y)(ψ(x)+O(ε))dz~,\displaystyle=\int_{{\mathbb{R}}^{2}}\text{\rm d}\widetilde{z}\int_{0}^{\frac{t}{\varepsilon}}\text{\rm d}u\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}y\text{\rm d}y^{\prime}p_{u}(y+\widetilde{z})p_{u}(y^{\prime}+\widetilde{z})f\left(y\right)f\left(y^{\prime}\right)\left(\psi(x)+O(\sqrt{\varepsilon})\right)\text{\rm d}\widetilde{z},

where O(ε)O(\sqrt{\varepsilon}) is uniformly dominated by CεC\sqrt{\varepsilon} for a constant C>0C>0. Moreover,

2dz~0tεdu2×2dydypu(y+z~)pu(y+z~)f(y)f(y)ψ(x)dz~\displaystyle\int_{{\mathbb{R}}^{2}}\text{\rm d}\widetilde{z}\int_{0}^{\frac{t}{\varepsilon}}\text{\rm d}u\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}y\text{\rm d}y^{\prime}p_{u}(y+\widetilde{z})p_{u}(y^{\prime}+\widetilde{z})f\left(y\right)f\left(y^{\prime}\right)\psi(x)\text{\rm d}\widetilde{z}
=ψ(x)0tεdu2×2dydyp2u(yy)f(y)f(y)\displaystyle=\psi(x)\int_{0}^{\frac{t}{\varepsilon}}\text{\rm d}u\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}y\text{\rm d}y^{\prime}p_{2u}(y-y^{\prime})f\left(y\right)f\left(y^{\prime}\right)
=ψ(x)4π2×2dydyf(y)f(y)ε|yy|24teuudu.\displaystyle=\frac{\psi(x)}{4\pi}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{2}}\text{\rm d}y\text{\rm d}y^{\prime}f\left(y\right)f\left(y^{\prime}\right)\int_{\frac{\varepsilon|y-y^{\prime}|^{2}}{4t}}^{\infty}\frac{e^{-u}}{u}\text{\rm d}u.

Then, (6.10) follows from l’Hôpital’s rule.

7 Peaks of 𝒵ϑ,ϕt(dx)\mathscr{Z}^{\vartheta,\phi}_{t}(\text{\rm d}x)

It is known that 𝒵ϑ,ϕt(dx)\mathscr{Z}^{\vartheta,\phi}_{t}(\text{\rm d}x) is singular with respect to Lebesgue measure [15, Theorem 10.5]. It follows from the fact that 1πδ2𝒵ϑ,ϕt(1B(x,δ)())\frac{1}{\pi\delta^{2}}\mathscr{Z}^{\vartheta,\phi}_{t}\left(1_{B(x,\delta)}(\cdot)\right) converges to 0 for Lebsegue a.e. x2x\in{\mathbb{R}}^{2} ([15, (10.9)]).

Thus, it follows that there exists (t,x)(0,)×2(t,x)\in(0,\infty)\times{\mathbb{R}}^{2} such that

1πδ2𝒵ϑ,ϕt(1B(x,δ)()) diverges as δ0.\displaystyle\frac{1}{\pi\delta^{2}}\mathscr{Z}^{\vartheta,\phi}_{t}\left(1_{B(x,\delta)}(\cdot)\right)\text{ diverges as $\delta\to 0$.}

Furthermore, [15, Theorem 10.6] says that for any t>0t>0 and ϑ\vartheta\in{\mathbb{R}} 𝒵tϑ,ϕ(dx)\mathscr{Z}_{t}^{\vartheta,\phi}(\text{\rm d}x) belongs to 𝒞0:=ε>0𝒞ε\mathcal{C}^{0-}:=\bigcap_{\varepsilon>0}\mathcal{C}^{-\varepsilon}, where 𝒞ε\mathcal{C}^{-\varepsilon} is the negative Besov-Hölder space of order ε-\varepsilonin the sense of [21, Definition 2.1].

In this section, we will give a “typical order of peak” of 𝒵ϑ,ϕt(dx)\mathscr{Z}^{\vartheta,\phi}_{t}\left(\text{\rm d}x\right).

Fix fCc+(2)f\in C_{c}^{+}({\mathbb{R}}^{2}) with 2f(x)dx=1\int_{{\mathbb{R}}^{2}}f(x)\text{\rm d}x=1. We define a random set

𝒯s,t(A,f,λ,ε):={(r,x):r[s,t],xA,𝒵rϑ,ϕ(fε(x))λlog1ε}for λ>0ε>0.\displaystyle\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon):=\left\{(r,x):r\in[s,t],x\in A,\mathscr{Z}_{r}^{\vartheta,\phi}(f_{\varepsilon}(\cdot-x))\geq\lambda\log\frac{1}{\varepsilon}\right\}\quad\text{for $\lambda>0$, $\varepsilon>0$}.

where A(2)A\subset\mathcal{B}({\mathbb{R}}^{2}) is a Borel set with finite Lebsegue measure and fε(x)=1εf(xε)f_{\varepsilon}(x)=\frac{1}{\varepsilon}f\left(\frac{x}{\sqrt{\varepsilon}}\right).

Theorem 7.1.

Fix ϑ\vartheta\in{\mathbb{R}} and ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}) with ϕ0\phi\not\equiv 0. Let A2A\subset{\mathbb{R}}^{2} be an open set and 0s<t<0\leq s<t<\infty.

  1. (1)

    We have

    lim¯λlim¯ε0𝒯s,t(A,f,λ,ε)𝒵ϑ,ϕu(fε(x))dxdu=0a.s.\displaystyle\varlimsup_{\lambda\to\infty}\varlimsup_{\varepsilon\to 0}\iint_{\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon)}\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon}(\cdot-x))\text{\rm d}x\text{\rm d}u=0\quad\text{a.s.}
  2. (2)

    There exists a non-random decreasing sequence {εn}n1\{\varepsilon_{n}\}_{n\geq 1} with εn0\varepsilon_{n}\to 0 such that

    P(λ>0lim¯n𝒯s,t(A,f,λ,εn) is not empty)>0.\displaystyle P\left(\bigcup_{\lambda>0}\varlimsup_{n\to\infty}\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon_{n})\text{ is not empty}\right)>0.

We can find that the peaks of 𝒵ϑ,ϕt(fε(x))\mathscr{Z}^{\vartheta,\phi}_{t}(f_{\varepsilon}(\cdot-x)) are of order log1ε\log\frac{1}{\varepsilon}. Thus, we may expect that 𝒵ϑ,ϕt(dx)\mathscr{Z}^{\vartheta,\phi}_{t}(\text{\rm d}x) belongs to the “logarithmic HaiHölder space” 𝒞0,1\mathscr{C}^{0,-1}, where we say ξ𝒮(2)\xi\in\mathscr{S}^{\prime}({\mathbb{R}}^{2}) belongs to 𝒞s,b\mathscr{C}^{s,b} for s<0s<0 and bb\in{\mathbb{R}} or s=0s=0 and b<0b<0 if ξ\xi belongs to the dual of 𝒞r\mathcal{C}^{r} with r={sif b0s+1if b<0r=\begin{cases}-\lfloor s\rfloor&\text{if }b\geq 0\\ -\lfloor s\rfloor+1\quad&\text{if }b<0\end{cases} and

ξs,b:=supε(0,1]supf𝒞r,suppfB(0,1),fγ1supx2|(1+log1ε)bεsξ,1ε2f(xε)|<.\displaystyle\|\xi\|_{s,b}:=\sup_{\varepsilon\in(0,1]}\sup_{\begin{smallmatrix}f\in\mathcal{C}^{r},\\ \mathrm{supp}f\subset B(0,1),\|f\|_{\gamma}\leq 1\end{smallmatrix}}\sup_{x\in{\mathbb{R}}^{2}}\left|\frac{(1+\log\frac{1}{\varepsilon})^{b}}{\varepsilon^{s}}\left\langle\xi,\frac{1}{\varepsilon^{2}}f\left(\frac{\cdot-x}{\varepsilon}\right)\right\rangle\right|<\infty.
Remark 7.2.

The reader may refer to [28, Definition 3.7] for the definition of 𝒞γ\mathscr{C}^{\gamma} and γ\|\cdot\|_{\gamma}. In [28], he referred the relationship between 𝒞α\mathscr{C}^{\alpha} and the Besov space Bα,B^{\alpha}_{\infty,\infty}. Then, 𝒞s,b\mathscr{C}^{s,b} is a slight modification of 𝒞α\mathscr{C}^{\alpha}. To our knowledge, there are no results concerning with the relationship between 𝒞s,b\mathscr{C}^{s,b} and the generalized Besov space (discussed in e.g. [37, 20, 1]).

We have the following result.

Theorem 7.3.

Fix ϑ\vartheta\in{\mathbb{R}}, and ϕCc+(2)\phi\in C_{c}^{+}({\mathbb{R}}^{2}) with ϕ0\phi\not\equiv 0. Then, {𝒵ϑ,ϕ(dx)}\left\{\mathscr{Z}_{\cdot}^{\vartheta,\phi}(\text{\rm d}x)\right\} is not a continuous 𝒞0,b\mathscr{C}^{0,b}-valued process with the uniform-on-compact topology for any b>1b>-1.

Proof of Theorem 7.3.

We can take fCc+(2)f\in C_{c}^{+}({\mathbb{R}}^{2}) in the statement in Theorem 7.3 with f𝒞γf\in\mathscr{C}^{\gamma}, supp(f)B(0,1)\mathrm{supp}(f)\subset B(0,1) and fγ1\|f\|_{\gamma}\leq 1.

Thus, (2) in Theorem 7.3 implies that for s=0s=0 and b>1b>-1,

sup0uTsupε(0,1]supf𝒞r,suppfB(0,1)supx2|(1+log1ε)b𝒵uϑ,ϕ(1ε2f(xε))|=\displaystyle\sup_{0\leq u\leq T}\sup_{\varepsilon\in(0,1]}\sup_{f\in\mathcal{C}^{r},\mathrm{supp}f\subset B(0,1)}\sup_{x\in{\mathbb{R}}^{2}}\left|{\left(1+\log\frac{1}{\varepsilon}\right)^{b}}\mathscr{Z}_{u}^{\vartheta,\phi}\left(\frac{1}{\varepsilon^{2}}f\left(\frac{\cdot-x}{\varepsilon}\right)\right)\right|=\infty

with positive probability for any T>0T>0. ∎

Proof of Theorem 7.1.

(1) Suppose that there exists c>0c>0 such that

lim¯λlim¯ε0𝒯s,t(A,f,λ,ε)𝒵ϑ,ϕu(fε(x))dxdu>0\displaystyle\varlimsup_{\lambda\to\infty}\varlimsup_{\varepsilon\to 0}\iint_{\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon)}\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon}(\cdot-x))\text{\rm d}x\text{\rm d}u>0

with positive probability.

Then, it is easy to see that

ϑ,ϕ(A)tϑ,ϕ(A)s\displaystyle\left\langle\mathscr{M}^{\vartheta,\phi}(A)\right\rangle_{t}-\left\langle\mathscr{M}^{\vartheta,\phi}(A)\right\rangle_{s} =limε04πlogεstA(𝒵ϑ,ϕu(fε(x)))2dxdu\displaystyle=-\lim_{\varepsilon\to 0}\frac{4\pi}{\log\varepsilon}\int_{s}^{t}\int_{A}\left(\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon}(\cdot-x))\right)^{2}\text{\rm d}x\text{\rm d}u
limε04πλ𝒯s,t(A,f,λ,ε)𝒵ϑ,ϕu(fε(x))dxdu\displaystyle\geq\lim_{\varepsilon\to 0}{4\pi\lambda}\iint_{\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon)}\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon}(\cdot-x))\text{\rm d}x\text{\rm d}u

for any λ>0\lambda>0 with positive probability and hence it is a contradiction.

(2)

Let B2B\subset{\mathbb{R}}^{2} be an open set with BB¯AB\subset\overline{B}\subset A. Let ψCb2(2)\psi\in C_{b}^{2}({\mathbb{R}}^{2}) be a positive function such that

0ψ(x)1for x2 and ψ(x)={0for xAc1for xB¯.\displaystyle 0\leq\psi(x)\leq 1\quad\text{for $x\in{\mathbb{R}}^{2}$ and }\psi(x)=\begin{cases}0\quad&\text{for $x\in A^{c}$}\\ 1\quad&\text{for $x\in\overline{B}$}\end{cases}.

We focus on the martingale tϑ,ϕ(ψ)\mathscr{M}_{t}^{\vartheta,\phi}(\psi). We can see

P(ϑ,ϕ(ψ)tϑ,ϕ(ψ)s>0)>0.\displaystyle P\left(\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{s}>0\right)>0.

Indeed, we have

P(ϑ,ϕ(ψ)tϑ,ϕ(ψ)s>0)14E[ϑ,ϕ(ψ)tϑ,ϕ(ψ)s]2E[(ϑ,ϕ(ψ)tϑ,ϕ(ψ)s)2]\displaystyle P\left(\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{s}>0\right)\geq\frac{1}{4}\frac{E\left[\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{s}\right]^{2}}{E\left[\left(\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{s}\right)^{2}\right]}

from the Paley-Zygmund inequality. Moreover, it follows from Lemma 3.14, (3.13), and Fatous’s lemma that the dominator in the right-hand side is bounded. Also, we can find from Corollary 3.13 that the numerator is strictly positive.

Also, we can see from Remark 6.6 that

ϑ,ϕ(ψ)tϑ,ϕ(ψ)s\displaystyle\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{s} =limn4πlogεnstA(𝒵ϑ,ϕu(fεn(x)))2ψ(x)2dxdu,a.s.,\displaystyle=-\lim_{n\to\infty}\frac{4\pi}{\log\varepsilon_{n}}\int_{s}^{t}\int_{A}\left(\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon_{n}}(\cdot-x))\right)^{2}\psi(x)^{2}\text{\rm d}x\text{\rm d}u,\quad\text{a.s.}, (7.1)

for a sequence {εn}n1\{\varepsilon_{n}\}_{n\geq 1} with εn0\varepsilon_{n}\to 0.

We set

Is,t,λ(1)(ε):=4πlogε𝒯s,t(A,f,λ,ε)(𝒵ϑ,ϕu(fε(x)))2ψ(x)2dxdu\displaystyle I_{s,t,\lambda}^{(1)}(\varepsilon):=-\frac{4\pi}{\log\varepsilon}\iint_{\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon)}\left(\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon}(\cdot-x))\right)^{2}\psi(x)^{2}\text{\rm d}x\text{\rm d}u
Is,t,λ(2)(ε):=4πlogε[s,t]×A\𝒯s,t(A,f,λ,ε)(𝒵ϑ,ϕu(fε(x)))2ψ(x)2dxdu.\displaystyle I_{s,t,\lambda}^{(2)}(\varepsilon):=-\frac{4\pi}{\log\varepsilon}\iint_{[s,t]\times A\backslash\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon)}\left(\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon}(\cdot-x))\right)^{2}\psi(x)^{2}\text{\rm d}x\text{\rm d}u.

We remark that if lim¯nIs,t,λ(1)(εn)>0\displaystyle\varliminf_{n\to\infty}I_{s,t,\lambda}^{(1)}(\varepsilon_{n})>0 for some λ>0\lambda>0, then lim¯n𝒯s,t(A,f,λ,εn)\varlimsup_{n\to\infty}\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon_{n}) is not empty set.

Thus, it is enough to prove that

:={ϑ,ϕ(ψ)tϑ,ϕ(ψ)s>0}a.s.m1{lim¯nIs,t,λm(1)(εn)>0}\displaystyle\mathcal{E}:=\left\{\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{s}>0\right\}\overset{\text{a.s.}}{\subset}\bigcup_{m\geq 1}\left\{\varliminf_{n\to\infty}I_{s,t,\lambda_{m}}^{(1)}(\varepsilon_{n})>0\right\}

for λm=1m\lambda_{m}=\frac{1}{m}.

We retake {εn}n1\{\varepsilon_{n}\}_{n\geq 1} with εn0\varepsilon_{n}\to 0 such tht in addition to (7.1),

st𝒵ϑ,ϕu(ψ)du=limεn0st2𝒵ϑ,ϕu(fεn(x))ψ(x)dxdu,a.s.\displaystyle\int_{s}^{t}\mathscr{Z}^{\vartheta,\phi}_{u}(\psi)\text{\rm d}u=\lim_{\varepsilon_{n}\to 0}\int_{s}^{t}\int_{{\mathbb{R}}^{2}}\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon_{n}}(\cdot-x))\psi(x)\text{\rm d}x\text{\rm d}u,\quad\text{a.s.}

We set

Js,t,λ(1)(ε):=𝒯s,t(A,f,λ,ε)𝒵ϑ,ϕu(fε(x))ψ(x)dxdu\displaystyle J_{s,t,\lambda}^{(1)}(\varepsilon):=\iint_{\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon)}\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon}(\cdot-x))\psi(x)\text{\rm d}x\text{\rm d}u
Js,t,λ(2)(ε):=[s,t]×A\𝒯s,t(A,f,λ,ε)𝒵ϑ,ϕu(fε(x))ψ(x)dxdu.\displaystyle J_{s,t,\lambda}^{(2)}(\varepsilon):=\iint_{[s,t]\times A\backslash\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon)}\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon}(\cdot-x))\psi(x)\text{\rm d}x\text{\rm d}u.

Since we have 𝒵uϑ,ϕ(fε(x))logελ-\frac{\mathscr{Z}_{u}^{\vartheta,\phi}(f_{\varepsilon}(\cdot-x))}{\log\varepsilon}\leq\lambda for (u,x)[s,t]×A\s,t(A,f,λ,ε)(u,x)\in[s,t]\times A\backslash\mathcal{F}_{s,t}(A,f,\lambda,\varepsilon),

Is,t,λ(2)(εn)\displaystyle I_{s,t,\lambda}^{(2)}(\varepsilon_{n}) =4πlogε[s,t]×A\𝒯s,t(A,f,λ,ε)(𝒵ϑ,ϕu(fε(x)))2ψ(x)2dxdu\displaystyle=-\frac{4\pi}{\log\varepsilon}\iint_{[s,t]\times A\backslash\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon)}\left(\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon}(\cdot-x))\right)^{2}\psi(x)^{2}\text{\rm d}x\text{\rm d}u
4πλ[s,t]×A\𝒯s,t(A,f,λ,ε)𝒵ϑ,ϕu(fε(x))ψ(x)dxdu=4πλJs,t,λ(2)(εn).\displaystyle\leq 4\pi\lambda\iint_{[s,t]\times A\backslash\mathcal{T}_{s,t}(A,f,\lambda,\varepsilon)}\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon}(\cdot-x))\psi(x)\text{\rm d}x\text{\rm d}u=4\pi\lambda J_{s,t,\lambda}^{(2)}(\varepsilon_{n}).

Since the right-hand side is bounded for {εn}\{\varepsilon_{n}\} and λ>0\lambda>0, there exists a random λ>0\lambda>0 such that

lim¯nIs,t,λ(2)(εn)12(ϑ,ϕ(ψ)tϑ,ϕ(ψ)s),\displaystyle\varlimsup_{n\to\infty}I_{s,t,\lambda}^{(2)}(\varepsilon_{n})\leq\frac{1}{2}\left(\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{s}\right),

on \mathcal{E}, and hence, it follows that

lim¯nIs,t,λ(1)(εn)12(ϑ,ϕ(ψ)tϑ,ϕ(ψ)s).\displaystyle\varliminf_{n\to\infty}I_{s,t,\lambda}^{(1)}(\varepsilon_{n})\geq\frac{1}{2}\left(\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{s}\right).

It is natural to expect that the values of 𝒵tϑ,ϕ(fε())\mathscr{Z}_{t}^{\vartheta,\phi}\left(f_{\varepsilon}(\cdot-*)\right) of order logε-\log\varepsilon contributes to 𝒵tϑ,ϕ(ψ)\mathscr{Z}_{t}^{\vartheta,\phi}(\psi) and ϑ,ϕ(ψ)t\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}. That is, we may expect

st𝒵ϑ,ϕu(ψ)dulimns,t(2,f,λ,εn)cs,t(2,f,1λ,εn)𝒵ϑ,ϕu(fεn(x)ψ(x)dxdu>0\displaystyle\int_{s}^{t}\mathscr{Z}^{\vartheta,\phi}_{u}(\psi)\text{\rm d}u\approx\lim_{n\to\infty}\iint_{\mathcal{F}_{s,t}({\mathbb{R}}^{2},f,\lambda,\varepsilon_{n})^{c}\cap\mathcal{F}_{s,t}({\mathbb{R}}^{2},f,\frac{1}{\lambda},\varepsilon_{n})}\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon_{n}}(\cdot-x)\psi(x)\text{\rm d}x\text{\rm d}u>0
ϑ,ϕ(ψ)tϑ,ϕ(ψ)slimn4πlogεn𝒯s,t(λ,εn)c𝒯s,t(2,f,1λ,εn)𝒵ϑ,ϕu(fεn(x))2ψ(x)2dxdu>0\displaystyle\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{t}-\left\langle\mathscr{M}^{\vartheta,\phi}(\psi)\right\rangle_{s}\approx-\lim_{n\to\infty}\frac{4\pi}{\log\varepsilon_{n}}\iint_{\mathcal{T}_{s,t}(\lambda,\varepsilon_{n})^{c}\cap\mathcal{T}_{s,t}({\mathbb{R}}^{2},f,\frac{1}{\lambda},\varepsilon_{n})}\mathscr{Z}^{\vartheta,\phi}_{u}(f_{\varepsilon_{n}}(\cdot-x))^{2}\psi(x)^{2}\text{\rm d}x\text{\rm d}u>0

for λ\lambda large enough. Let

Ls,t(f,ε,λ):=|𝒯s,t(2,f,λ,εn)c𝒯s,t(2,f,1λ,εn)|.\displaystyle L_{s,t}(f,\varepsilon,\lambda):=\left|\mathcal{T}_{s,t}({\mathbb{R}}^{2},f,\lambda,\varepsilon_{n})^{c}\cap\mathcal{T}_{s,t}\left({\mathbb{R}}^{2},f,\frac{1}{\lambda},\varepsilon_{n}\right)\right|.

Then, we would find that

Ls,t(f,ε,λ)1log1ε\displaystyle L_{s,t}(f,\varepsilon,\lambda)\approx\frac{1}{\log\frac{1}{\varepsilon}}

for large λ>0\lambda>0.

Acknowledgemments This work was supported by JSPS KAKENHI Grant Number JP22K03351, JP23K22399. The author thanks Prof. Nikos Zygouras for useful comments.

References

  • Alm [05] Alexandre Almeida. Wavelet bases in generalized Besov spaces. J. Math. Anal. Appl., 304(1):198–211, 2005.
  • BC [98] Lorenzo Bertini and Nicoletta Cancrini. The two-dimensional stochastic heat equation: renormalizing a multiplicative noise. Journal of Physics A: Mathematical and General, 31(2):615, 1998.
  • BG [97] Lorenzo Bertini and Giambattista Giacomin. Stochastic Burgers and KPZ equations from particle systems. Communications in mathematical physics, 183(3):571–607, 1997.
  • Bur [73] D. L. Burkholder. Distribution function inequalities for martingales. Ann. Probability, 1:19–42, 1973.
  • CC [22] Francesco Caravenna and Francesca Cottini. Gaussian limits for subcritical chaos. Electron. J. Probab., 27:Paper No. 81, 35, 2022.
  • CCM [20] Francis Comets, Clément Cosco, and Chiranjib Mukherjee. Renormalizing the Kardar-Parisi-Zhang equation in d3d\geq 3 in weak disorder. J. Stat. Phys., 179(3):713–728, 2020.
  • CCM [24] Francis Comets, Clément Cosco, and Chiranjib Mukherjee. Space-time fluctuation of the Kardar-Parisi-Zhang equation in d3d\geq 3 and the Gaussian free field. Ann. Inst. Henri Poincaré Probab. Stat., 60(1):82–112, 2024.
  • CD [20] Sourav Chatterjee and Alexander Dunlap. Constructing a solution of the (2+1)(2+1)-dimensional KPZ equation. Ann. Probab., 48(2):1014–1055, 2020.
  • CNN [22] Clément Cosco, Shuta Nakajima, and Makoto Nakashima. Law of large numbers and fluctuations in the sub-critical and L2L^{2} regions for SHE and KPZ equation in dimension d3d\geq 3. Stochastic Process. Appl., 151:127–173, 2022.
  • CSZ [17] Francesco Caravenna, Rongfeng Sun, and Nikos Zygouras. Universality in marginally relevant disordered systems. Ann. Appl. Probab., 27(5):3050–3112, 2017.
  • [11] Francesco Caravenna, Rongfeng Sun, and Nikos Zygouras. The Dickman subordinator, renewal theorems, and disordered systems. Electron. J. Probab., 24:Paper No. 101, 40, 2019.
  • [12] Francesco Caravenna, Rongfeng Sun, and Nikos Zygouras. On the moments of the (2+1)(2+1)-dimensional directed polymer and stochastic heat equation in the critical window. Comm. Math. Phys., 372(2):385–440, 2019.
  • CSZ [20] Francesco Caravenna, Rongfeng Sun, and Nikos Zygouras. The two-dimensional KPZ equation in the entire subcritical regime. Annals of Probability, 48(3):1086–1127, 2020.
  • CSZ [23] Francesco Caravenna, Rongfeng Sun, and Nikos Zygouras. The critical 2d Stochastic Heat Flow. Invent. Math., 233(1):325–460, 2023.
  • CSZ [24] Francesco Caravenna, Rongfeng Sun, and Nikos Zygouras. The critical 2d stochastic heat flow and related models, 2024.
  • DF [94] Donald A. Dawson and Klaus Fleischmann. A super-Brownian motion with a single point catalyst. Stochastic Process. Appl., 49(1):3–40, 1994.
  • DGRZ [20] Alexander Dunlap, Yu Gu, Lenya Ryzhik, and Ofer Zeitouni. Fluctuations of the solutions to the KPZ equation in dimensions three and higher. Probab. Theory Related Fields, 176(3-4):1217–1258, 2020.
  • EK [86] Stewart N. Ethier and Thomas G. Kurtz. Markov processes. Wiley Series in Probability and Mathematical Statistics: Probability and Mathematical Statistics. John Wiley & Sons, Inc., New York, 1986. Characterization and convergence.
  • FH [14] Peter K. Friz and Martin Hairer. A course on rough paths. Universitext. Springer, Cham, 2014. With an introduction to regularity structures.
  • FL [06] Walter Farkas and Hans-Gerd Leopold. Characterisations of function spaces of generalised smoothness. Ann. Mat. Pura Appl. (4), 185(1):1–62, 2006.
  • FM [17] Marco Furlan and Jean-Christophe Mourrat. A tightness criterion for random fields, with application to the Ising model. Electron. J. Probab., 22:Paper No. 97, 29, 2017.
  • FV [10] Peter K. Friz and Nicolas B. Victoir. Multidimensional stochastic processes as rough paths, volume 120 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2010. Theory and applications.
  • GIP [15] Massimiliano Gubinelli, Peter Imkeller, and Nicolas Perkowski. Paracontrolled distributions and singular PDEs. Forum Math. Pi, 3:e6, 75, 2015.
  • GJ [14] Patrícia Gonçalves and Milton Jara. Nonlinear fluctuations of weakly asymmetric interacting particle systems. Arch. Ration. Mech. Anal., 212(2):597–644, 2014.
  • GQT [21] Yu Gu, Jeremy Quastel, and Li-Cheng Tsai. Moments of the 2D SHE at criticality. Probab. Math. Phys., 2(1):179–219, 2021.
  • GRZ [18] Yu Gu, Lenya Ryzhik, and Ofer Zeitouni. The Edwards-Wilkinson limit of the random heat equation in dimensions three and higher. Comm. Math. Phys., 363(2):351–388, 2018.
  • [27] M. Hairer. Solving the KPZ equation. In XVIIth International Congress on Mathematical Physics, page 419. World Sci. Publ., Hackensack, NJ, 2014.
  • [28] M. Hairer. A theory of regularity structures. Invent. Math., 198(2):269–504, 2014.
  • JN [24] Stefan Junk and Shuta Nakajima. Equivalence of fluctuations of discretized she and kpz equations in the subcritical weak disorder regime, 2024.
  • JS [87] Jean Jacod and Albert N Shiryaev. Limit theorems for stochastic processes. Springer, 1987.
  • Jun [23] Stefan Junk. Fluctuations of partition functions of directed polymers in weak disorder beyond the l2l^{2}-phase, 2023.
  • Jun [24] Stefan Junk. Local limit theorem for directed polymers beyond the l2l^{2}-phase, 2024.
  • KPZ [86] Mehran Kardar, Giorgio Parisi, and Yi-Cheng Zhang. Dynamic scaling of growing interfaces. Physical Review Letters, 56(9):889, 1986.
  • Kup [16] Antti Kupiainen. Renormalization group and stochastic PDEs. Ann. Henri Poincaré, 17(3):497–535, 2016.
  • LL [10] Gregory F. Lawler and Vlada Limic. Random walk: a modern introduction, volume 123 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2010.
  • LZ [22] Dimitris Lygkonis and Nikos Zygouras. Edwards-Wilkinson fluctuations for the directed polymer in the full L2L^{2}-regime for dimensions d3d\geq 3. Ann. Inst. Henri Poincaré Probab. Stat., 58(1):65–104, 2022.
  • Mou [01] Susana Moura. Function spaces of generalised smoothness. Dissertationes Math. (Rozprawy Mat.), 398:88, 2001.
  • MSZ [16] Chiranjib Mukherjee, Alexander Shamov, and Ofer Zeitouni. Weak and strong disorder for the stochastic heat equation and continuous directed polymers in d3d\geq 3. Electron. Commun. Probab., 21:Paper No. 61, 12, 2016.
  • MU [18] Jacques Magnen and Jérémie Unterberger. The scaling limit of the KPZ equation in space dimension 3 and higher. J. Stat. Phys., 171(4):543–598, 2018.
  • Mue [91] Carl Mueller. On the support of solutions to the heat equation with noise. Stochastics Stochastics Rep., 37(4):225–245, 1991.
  • Per [02] Edwin Perkins. Dawson-Watanabe superprocesses and measure-valued diffusions. In Lectures on probability theory and statistics (Saint-Flour, 1999), volume 1781 of Lecture Notes in Math., pages 125–324. Springer, Berlin, 2002.
  • Spi [76] Frank Spitzer. Principles of random walk, volume Vol. 34 of Graduate Texts in Mathematics. Springer-Verlag, New York-Heidelberg, second edition, 1976.
  • Tao [11] Terence Tao. An introduction to measure theory, volume 126 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2011.
  • Tsa [24] Li-Cheng Tsai. Stochastic heat flow by moments, 2024.
  • Wal [86] John B. Walsh. An introduction to stochastic partial differential equations. In École d’été de probabilités de Saint-Flour, XIV—1984, volume 1180 of Lecture Notes in Math., pages 265–439. Springer, Berlin, 1986.